You are on page 1of 21

REACTORS, KINETICS, AND CATALYSIS

JOURNAL REVIEW

Monoliths as Multiphase Reactors: A Review


Shaibal Roy, Tobias Bauer, and Muthanna Al-Dahhan
Chemical Reaction Engineering Laboratory, Washington University, St. Louis, MO 63130

Peter Lehner and Thomas Turek


Bayer Technology Services, Bayer AG, D-51368, Leverkusen, Germany

DOI 10.1002/aic.10268
Published online in Wiley InterScience (www.interscience.wiley.com).

Monolith reactors are being studied as a replacement for conventional multiphase


reactors such as trickle-bed reactors, slurry reactors, and slurry bubble column reactors
for gas–liquid–solid reactions. Reactors with monolith catalyst packing have been found
to be hydrodynamically superior to existing industrial reactors. This review covers
multiphase reactions carried out in monolith reactors by various researchers. It first
defines the monolith reactor and looks into the geometrical aspects of monolith. The
section dealing with hydrodynamics reviews pressure drop, phase holdup, flow distribu-
tion, and dispersion characteristics. This study also considers the tools used to charac-
terize the hydrodynamic parameters and their typical values. Although the available
literature is scarce, monoliths are considered to have superior mass transfer character-
istics. This review lists the mass transfer correlations for each category (gas–liquid,
gas–solid, liquid–solid). The last section discusses the reaction aspects of monolith
reactors. The ultimate goal is to implement such reactors for multiphase reactions. This
section also compares the performance of monolith reactors with conventional multiphase
reactors and lists the various reactor models reported to predict the overall performance
of monolith reactors. © 2004 American Institute of Chemical Engineers AIChE J, 50: 2918 –2938,
2004
Keywords: monolith, structured packing, hydrodynamics, Review

Introduction Different types of reactors have been used for three-phase


Among the various chemical reactions occurring in broad gas–liquid–solid reaction applications. The major ones are the
range of industrial application areas, catalytic gas–liquid–solid stirred tank slurry reactor, the slurry bubble column reactor,
and the packed-bed reactor. The choice is governed by the
reactions are widespread. These reactions occur extensively in
reaction chemistry, the ease of use and manufacture of reactor
chemical, petroleum, petrochemical, biochemical, material,
types, and the enhancement of mass transfer for mass transfer
and environmental industrial processes for a wide variety of
limited reactions. Each reactor type has its own advantages and
products (such as hydrogenation, oxidation, and alkylation).
drawbacks. Slurry catalysts are very small (micrometer scale,
Two notable examples are the catalytic hydrogenation of pe-
5–50 ␮m), whereas packed-bed particles are larger (millimeter
troleum fractions to remove sulfur impurities and the catalytic
scale, 1–3 mm). In general it can be stated that larger particles
oxidation of liquid hydrocarbons with air or oxygen (Leven-
are kinetically less efficient because of intraparticle diffusion
spiel, 1996).
limitation. The solid–liquid slurry and the slurry bubble col-
umn reactors offer very simple reactor geometry, high heat
Correspondence concerning this article should be addressed to M. Al-Dahhan at removal, excellent mass transfer characteristics, and a high
muthanna@che.wustl.edu.
effectiveness factor because of the very small particle size. The
© 2004 American Institute of Chemical Engineers solid catalyst is suspended in the liquid medium through which

2918 November 2004 Vol. 50, No. 11 AIChE Journal


gas is dispersed. The major disadvantages of these reactors are cause redistribution of gas–liquid flow over the reactor length
the separation of product and catalyst, and catalyst attrition. is not possible.
The filtration of fine particles, needed to separate the catalyst This work reviews recent advances made in the use of
from liquid product, often makes this an inconvenient reactor monoliths as multiphase flow reactors. The review starts by
type. A packed-bed reactor, such as the trickle-bed reactor defining the monolith structure, and then explores various
(TBR), is much more convenient, although larger particles hydrodynamic parameters, such as flow regime, liquid holdup,
must be used to guarantee moderate pressure drop. In these pressure drop, and dispersion. This is followed by a section
reactors, gas and liquid flow cocurrently downward over the dealing with mass transfer characteristics. These parameters
catalyst. Liquid flowing in a bed tends to form channels and are needed to predict the overall performance, design, and
bypasses—maldistribution of reactants is unavoidable. On the scale-up of monolith reactor systems. The overall performance
catalyst surface, where the liquid is either depleted or imper- is most frequently based on the intrinsic rate of model reactions
fectly covers the catalyst surface, dry areas are encountered: on monolith catalyst; thus, various kinetic measurements per-
these substantially reduce the liquid–solid contacting efficiency formed on such catalyst supports at an intrinsic rate level are
of the trickle-bed reactor. Moreover, local hot spots may de- reported as well. Finally, the section on reactor performance
velop and cause runaways. Adding to the problem are the low reviews the performance of monolith reactor systems, combin-
gas–liquid velocities required to avoid excessive pressure drop. ing the hydrodynamic and kinetic parameters.
This requirement results in high operational costs and low We attempt to summarize what is known about hydrody-
productivity. Another major drawback of conventional reactors namics, transport phenomena, and kinetics in monolith reactors
for multiphase reactions is the difficulty of scale-up to indus- for three-phase applications. It is hoped that this will stimulate
trial size units (Kapteijn et al., 2001). additional research activities, which is need to further advance
To overcome the above-mentioned difficulties, research has our understanding of this reactor type.
led to the use of structured packing instead of a random packed
bed. Monolith Structure
There are different types of structured packing: packing
made of regular ceramic or metal support is called “monolith,” The pace of research on monoliths as multiphase reactions
whereas other types include “sandwich packing” and “open has picked up in the last decade. Although most of the studies
cross flow structures.” The characteristic features of these types are presently confined to the laboratory and pilot scale, work on
of packing are uniform flow distribution, low pressure drop, the hydrogenation of alkylanthraquinones in hydrogen perox-
and enhanced mass transfer. The main advantage, and the ide production has reached full-scale operation (Albers et al.,
explanation of its noticeable performance enhancement, lies in 2001; Schöön, 1989). This section describes the characteristic
the manner how gas, liquid, and catalyst contact. As a result of features of the monolith structure for the use as multiphase
the application of regular structure, the scale-up to industrial reactor packing at the laboratory scale.
Monolith structures are essentially a single structure with
relevant size is considerably easier.
many thin, vertical, parallel channels, separated from each
Through the application of these new structures, traditional
other by walls. The channels are usually rectangular, but tri-
unit operations and reactors can be replaced by new, highly
angles, hexagons, or more complex geometries also exist. To
productive and energy-efficient reactors. The new reactors will
increase the surface area, internal fins can also be provided.
provide more process security, less by-product, and reduced
Furthermore, the fins have a stabilizing effect on the gas–liquid
dimensions for the same productivity. Stankiewicz and Mou-
flow and allow operation in countercurrent mode without
lijin (2000) described these novel developments, which have
flooding (Lebens et al., 1997, 1999a,b,c). To enhance turbu-
led to dramatic improvements in equipment and methods, such
lence inside monolith channels, patented monolith structures
as process intensification. have also been developed in which either corrugation has been
This article will consider only the use of monolithic catalyst provided or channels have been interconnected for radial trans-
carrier for gas–liquid–solid reaction systems. port (Cybulski et al., 1998). Obviously, hydrodynamics of such
Monoliths have been successfully used for the abatement of specialty monoliths will be different from that of the straight-
NOx and CO emissions from automobile engines and in mass channel monoliths. Such details are currently unavailable in the
transfer operations such as distillation and absorptions (Cybul- open literature.
ski et al., 1998; Ellenberger et al., 1999). The number of channels per cross section, the “cell density,”
However, their potential for use in multiphase reactions has typically ranges between 100 and 1200 channels per square
not yet been fully realized. For slurry, trickle-bed, and mono- inch (cpsi). The void fraction varies between 0.5 and 0.9 and is
lith reactors, a number of experimental and theoretical com- frequently expressed as the open frontal area (OFA). Typical
parisons have been made between typical parameters of cata- values for the wall thickness range between 0.006 and 0.05 cm.
lytic gas–liquid reactions (see, for example, Mazzaroni et al., A monolith structure is characterized by the wall thickness and
1987; Nijhuis et al., 2001; Stankiewicz et al., 2001). These cell density, which are independent of each other. Various
studies clearly show the superior behavior of the monolithic geometric parameters related to a monolith having a square
reactor over traditional three-phase reactors with respect to cross section are listed below and correspond to Figure 1.
productivity and selectivity. Currently, Akzo Nobel produces Cell Density
hydrogen peroxide on a large scale using a monolith catalyst
for the hydrogenation of anthraquinone to the corresponding
hydroquinones (Albers et al., 2001). However, for successful 1
n⫽ (1)
application uniform gas–liquid distribution is inevitable be- L2

AIChE Journal November 2004 Vol. 50, No. 11 2919


Satterfield et al., 1977), and hole plates, glass frits, and static
mixers for upflow mode (Thulasidas et al., 1995b); still other
variations exist.
Several configurations of the monolith reactor setup are
found in the literature. In the “monolith froth reactor,” reactant
mixture in the form of a froth is passed upward in the reactor
(Crynes et al., 1995; Thulasidas et al., 1995b). In the “in-line
monolith reactor,” the monolith pieces are incorporated in the
pipes carrying reactants (Stankiewicz et al., 2001). In the
“mononolith loop reactor” (Heiszwolf et al., 2001a), and in the
“ejector-driven monolith loop reactor” (Broekhuis et al., 2001),
the gaseous reactant is sucked into the reactor by the liquid
flow, making a closed loop of gas flow. The basic element in
these reactors is the monolithic support; however, they vary
Figure 1. Cross-section of a single cell (not to scale).
with respect to cell density and void fraction. They also vary in
hydrodynamics, such as in the direction of flow and recircula-
Open Frontal Area tion of the liquid and the gas phases.
In general, in multiphase reactors it is preferable to have a
high mass transfer rate to the catalyst surface. This rate can be
共L ⫺ t w兲 2
OFA ⫽ n共L ⫺ t w兲 2 ⫽ (2) achieved by a high surface-to-volume ratio and a short diffu-
L2 sion length of gas through the liquid. The monolith structure
provides a very high ratio, especially for monoliths with high
Geometric Surface Area cell density. For example, a monolith with a cell density of 600
cpsi and an open frontal area of 82% has a surface-to-volume
共L ⫺ t w兲 ratio of 3476 m2/m3 (data provided by Corning, Inc.). A short
GSA ⫽ 4n共L ⫺ t w兲 ⫽ 4 (3) diffusion length can be obtained by operating the monolith in
L2
two-phase slug flow (Taylor flow) or in the annular flow
Hydraulic Diameter regime. Taylor flow is characterized by a train of liquid slugs
and gas bubbles moving consecutively upward or downward

冉 冊
through the channel. The gas bubble length is several times the
OFA 共L ⫺ t w兲 channel diameter, and the gas bubble diameter is almost equal
dh ⫽ 4 ⫽ (4)
GFA 4 to the channel diameter, so just a thin liquid film separates the
gas from the wall. Typically, the liquid film thickness is in the
The parameters in Eqs. 1–3 are defined as: t, wall thickness; L, range of 30 to 70 ␮m (Irandoust et al., 1989a; experiments in
length from one channel wall center to the other; R, fillet a circular glass capillary). The annular flow regime is charac-
radius. terized by the flow of liquid along the channel wall as a thin
Monoliths are industrially produced by extrusion of a paste film and gas flow in the core of the channel. This regime can be
containing catalyst particles or by extrusion of a support on achieved for low liquid and high gas flow rates.
which the catalyst can be coated (washcoating). A general In summary, the primary advantages of the monolith packing
overview about characteristics, fabrication, and typical appli- are low pressure drop in the channels and high mass transfer
cations can be found in Williams (2001), Garcia-Bordeje et al. rates attributed to small diffusion paths. Therefore, carrying out
(2002), and Gulati (1998). A previous work by Lachman and equilibrium-limited reactions, such as hydrogenation and par-
Williams (1992) also gives a good introduction to monolith tial oxidation in monoliths, is a promising proposition. More-
production, raw materials, and its application in catalytic two- over, many hydrogenation reactions of industrial relevance are
phase processes. consecutive reactions, and for such reactions, monoliths have
Inside the monolith, the channels are separated by the chan- been shown to perform better with respect to productivity and
nel walls, and therefore no radial mixing occurs. This case is selectivity (Nijhuis et al., 2001).
similar to an ideal plug-flow reactor, and therefore the reactor
yield will be high.
The operating mode and flow distribution have strong effects Hydrodynamics
on the performance of these reactors. The reactor can be One of the most important criteria for selection of a mul-
operated in batch and in continuous modes. In the batch mode, tiphase reactor is choosing appropriate hydrodynamics
liquid is continuously circulated through the monolith unit until (Krishna et al., 1994). Apart from high conversion, factors such
the desired conversion is met. In the continuous mode, the as low pressure drop, high operating velocities without flood-
liquid flows only once through the monolith core. There are ing and instabilities, the extent of axial mixing, and the proper
three different flow arrangements possible: cocurrent down- distribution of reactant for effective utilization of catalyst are
flow, cocurrent upflow, and countercurrent flow. Different part of the wish list for reactor design. A basic understanding
types of distributors can be used to achieve the desired uniform of hydrodynamics of monolith reactors is essential to their
flow distribution over the monolith cross section. Literature design, scale-up, scale-down, and performance.
cites the use of spray nozzles, shower heads, and ejectors for Researchers in this field range from experiments to develop
downflow (Broekhuis et al., 2001; Hatziantoniou et al., 1984; models for predicting the results a priori. Experimental studies

2920 November 2004 Vol. 50, No. 11 AIChE Journal


channels, it is virtually impossible to determine the flow char-
acteristics and flow regime changes inside the monoliths.
Therefore, new noninvasive measurement techniques, such as
magnetic resonance imaging (MRI), can aid in depiction of
flow regimes inside monoliths. Two recent reports on the use of
MRI to determine slug flow were published by Mantle et al.
(2002) and Gladden et al. (2003). The extension of the appli-
Figure 2. A monolith channel operated in slug-flow re- cation to other flow regimes is desirable. Table 1 summarizes
gime (from Coleman et al., 1999). the work performed in single capillary tubes and monoliths.
The choice of the particular regime for industrial applica-
tions depends on the nature of reaction and the process condi-
have primarily been conducted in single circular capillary
tions. Flow regimes are mainly influenced by the gas and liquid
tubes, for ease of visualization, but have also been conducted in
properties, superficial velocities, and the diameter of the cap-
laboratory-scale monoliths (1–5 cm diameter). Researchers
illary and monolith channel.
have developed both simple semiempirical correlations as well
The overall gas–liquid holdup, pressure drop, and distribu-
as detailed numerical simulations, taking into account all the
tion inside the reactor depend on the flow regime in which the
possible physics.
reactor is working.
The hydrodynamic parameters such as flow regime, two-
Most of the authors have described the flow regimes as
phase flow pressure drop, and liquid-holdup mainly depend on
consisting of bubbly flow at low gas and high liquid velocities.
superficial liquid and gas velocities, physical properties of the
They have found slug flow, churn flow, and annular flow at
fluids, as well as on the type of structured packing used, such
high gas and low liquid flows, as shown in Table 1.
as random packing, monolith, static mixer, or any other struc-
Although the above regimes were observed in cocurrent
ture. Furthermore, it also depends on the mode of operation:
upflow, any difference in downflow regimes caused by the
whether it is operated cocurrently or countercurrently, upflow
difference in flow direction is small because of the relatively
or downflow, and so forth.
negligible influence of gravitational force in capillary tubes
This section reviews the following: the flow regime normally
(Cybulski et al., 1998).
encountered in small-diameter capillary tubes as well as in
In the real world of monoliths, it remains to be seen how
monoliths for a range of gas and liquid velocities; studies on
capillary analysis can be extended to monolith beds, consider-
pressure drop in monoliths and delineate the relevant correla-
ing the fact that the monolith channels are mostly noncircular
tions; and the gas–liquid holdup (most often, though, the pres-
in cross section and the surface roughness properties could be
sure drop studies are related to holdup calculations and axial
significantly different from those of glass tubes. It is question-
dispersion). A major concern of researchers dealing with mul-
able whether the exiting theories obtained from experiments in
tiphase reactors is the gas–liquid distribution inside the reactor
capillary tubes can be extended to monoliths. Further investi-
and the means to achieve uniform distribution. There is a
gations into the effects of capillary forces, caused by sharp
general disagreement among researchers regarding the degree
corners of noncircular channels on the two-phase flow pattern,
of axial and radial dispersion in monoliths. Some of the studies
are inevitable. The authors believe that only noninvasive mea-
reviewed in this section have found the dispersion to be sig-
surement techniques, such as MRI in conjunction with the
nificant.
consideration of the gas/liquid distribution device, can provide
Flow regime
A number of flow regimes have been identified in capillary
channels. These regimes are dispersed bubbly flow, slug flow
(Taylor flow) (see Figure 2), churn flow, and annular flow.
Primarily, these studies have been carried out by researchers
studying heat exchangers or in the field of nuclear engineering
(Coleman et al., 1999; Mishima et al., 1996; Triplett et al.,
1999a,b; Zhao et al., 2001). The common aim of these re-
searchers is to establish maps to visualize the flow pattern
inside the capillaries and to define the transition areas and flow
pattern maps using gas and liquid superficial velocities as
coordinates.
In a more recent study, Mishima et al. (1996) performed
experiments in capillary tubes. Figure 3 shows an experimen-
tally developed flow regime map for a glass capillary of 2.05
mm diameter, in which the model correlations developed for
large diameter tubes have been superimposed. Similar flow
maps from 1.05-, 3.12-, and 4.08-mm tubes are also available
Figure 3. Flow regime map in a single capillary (d ⴝ
(Mishima et al., 1996).
2.05 mm).
In the world of monoliths, similar flow regimes exist but
The open symbols indicate clearly identified flow regimes and
they are difficult to visualize. Because of the opaque nature of the solid symbols fall in the gray areas (from Mishima et al.,
the monolith material and the small hydraulic diameter of the 1996).

AIChE Journal November 2004 Vol. 50, No. 11 2921


Table 1. Summary of Flow Regime Studies
Reference Structure and Flow Flow Range Reactor Dimensions Flow Regimes Studied
Scatterfield et al., Monolith with three Air–water L R ⫽ 122 cm Annular flow, slug flow
1977 different
distributors
Air–cyclohexane D R ⫽ 2.54 cm
u L ⫽ 0.33 to 6.58 cm/s n ⫽ 200 or 360 cpsi
Cocurrent downward u G ⫽ 0–150 cm/s Block length ⫽ 7.6 or
15.2 cm
Irandoust et al., 1989a Capillary glass tube Liquid loading ⫽ 0.04 L R ⫽ 200 mm Slug flow
to 0.43 m3/m2 s
Cocurrent downward u tot ⫽ 0.27 to 0.53 m/s d c ⫽ 1.5 mm
Mishima et al., 1996 Capillary tube Air–water d c ⫽ 1.05–4.08 mm Bubbly flow, slug flow,
churn flow, annular
flow
u L ⬃ 0.01–0.91 m/s
Cocurrent upflow u G ⬃ 0.09–80 m/s
(depending on d c )
Reineck et al., 1996 Monolith u L ⫽ 0–0.2 m/s LR ⫽ 2 m Bubbly flow, plug flow,
aerated slug flow,
annular flow
u G ⫽ 0–8 m/s D R ⫽ 118 mm
Cocurrent downward n ⫽ 300 and 400 cpsi
Lebens et al., 1997 Monoliths with n-Decane–air L R ⫽ 0.5 m Annular flow
internal fins and
beveled ends
u L ⫽ 0.009–0.05 m/s d c ⫽ 4.5 mm
Countercurrent u G ⫽ 0–5 m/s
Mewes et al., 1999 Monolith, TBR u L ⫽ 0.01–0.14 m/s L R ⫽ 30 mm Slug flow unaerated,
(without air) aerated
Cocurrent downward u L ⫽ 0.26–0.39 m/s D R ⫽ 120 mm
u G ⫽ 0–5 m/s d c ⫽ 2 mm
Coleman et al., 1999 Glass capillary u L ⫽ 0.01–10 m/s d h ⫽ 1.3 to 5.5 Bubbly, dispersed and
elongated bubbly
flow, slug flow,
wavy, annular wavy
and annular flow
Cocurrent vertical u G ⫽ 0.1–100 m/s
flow
Triplett et al., 1999a,b Glass capillary u L ⫽ 0.02–8 m/s d c ⫽ 1.1 and 1.45 Bubbly flow, churn
mm flow, slug flow,
annular flow
Cocurrent vertical u G ⫽ 0.02–80 m/s
flow
Zhao et al., 2001 Triangular glass u L ⫽ 0.1–10 m/s d h ⫽ 0.866, 1.443, Dispersed bubbly flow,
capillary and 2.886 mm slug flow, churn
flow, annular flow
Cocurrent upflow u G ⫽ 0.1–100 m/s

proper flow maps and transition criteria for monoliths. Unfor- It is important in determining the energy losses, the sizing of
tunately, such a technique is costly, limited to reactor scale, and the compression equipment, liquid holdup, gas–liquid interfa-
only a few research groups work on this topic. Furthermore, cial area, and mass transfer coefficient (Al-Dahhan et al.,
pressure drop measurements could give important information 1994). High pressure drop through the system not only requires
about the flow regime and flow regime transitions inside the high energy input to the system, it also prohibits the unit from
monolith bed. being operated at high gas and liquid velocities, and thus the
Moreover, the turbulent regime should be more thoroughly throughput is limited. The use of novel structured catalystic
investigated because it could significantly enhance the produc- packing like monoliths addresses these concerns and reduces
tivity at a pressure drop comparable to that of packed-bed investment and operation costs.
reactors. The factors influencing the instabilities in flow and Starting from Ergun (1954), the estimation of pressure drop
transition from one regime to another need to be thoroughly along packed beds has been well documented over the years.
investigated. Besides a number of empirical correlations (Pinna et al., 2001)
phenomenological models have also been developed (Al-Dah-
Pressure drop han et al., 1997).
Operations of multiphase processes, especially packed-bed As mentioned earlier, one of the major advantages of struc-
reactors, are always associated with pressure losses because of tured packing is low pressure drop along the bed, which en-
the inner design of the reactors. Pressure drop represents the ables the unit to run at higher capacity without encountering
energy dissipated caused by fluid flow through the reactor bed. hydrodynamic instability. Table 2 lists the pressure drop cor-

2922 November 2004 Vol. 50, No. 11 AIChE Journal


Table 2. Pressure Drop Equations and Correlations for Capillaries and Monoliths
Author Pressure Drop Equation and Correlation
Standard pressure drop expression Frictional pressure drop per unit length of single phase:

冉 冊
⌬P
L f,i
⫽ 2f
␳i ui2
dh
(5)

Friction factor (Darcy, Fanning):


fD fF
f⫽ ⫽ (6)
8 2
For laminar flow, Re ⬍ 2100 (Hagen–Poiseulle law):
64 16
fD ⫽ fF ⫽ (7)
Re Re

冉 冊
⌬P
⌬L i

32␮␳i
d2h
共i ⫽ G, L兲 (8)

Satterfield et al. (1977) ⌬P m ⫽ ⌬P f ⫹ ⌬P or ⫺ g ␳ L L␧ L (9)


2
N共Vorif ⫺ Vtot
2

⌬porif ⫽ 关␧L ␳L ⫹ 共1 ⫺ ␧L 兲␳G 兴 (10)
2
Grolman et al. (1996) ⌬P ⫽ ⌬P TP,f ⫹ ⌬P G (11)
64 L 1
⌬PTP,f ⫽ ␧L ␳ 共u ⫹ uG 兲2 (12)
ReTP dc 2 L L
⌬P G ⫽ Cu G with C ⫽ 45,000 Pa s⫺1 m (13)

Mewes et al. (1999) ⌬P uL uG ␳L ␧gG


⫽ ⫺␧L ␳L g ⫹ 32␮L 2 ⫹ 32␮G 2 ⫹ 共uL ⫹ uG 兲2 (14)
⌬L dC dC 2 Lb

␧gG 共1⫺␧ G 兲0.15


⫽ ␧ G ⱖ30% (15)
L b L b0 共1⫺0.15兲
␧gG ␧G
⫽ ␧G ⱕ 30% (16)
Lb Lb0

Heiszwolf et al. (2001a) Entry-region friction model:


fTP ReL ⫽ 共f Re兲L ␧L 1 ⫹ 0.065 冉 冊 册 Lb
dc ReL
⫺0.66
(17)

Heiszwolf et al. (2001a)


冉 冊
⌬P
⌬L
1
⫽ ⫺fTP ␳L uTP
2
2
4
d
⫹ ␧L ␳L g (18)

F ␳L uTP dh
fTP ⫽ ReTP ⫽
ReTP ␮L

F ⫽ 18 (200 cpsi, OFA ⫽ 0.74)


F ⫽ 22 (400 cpsi, OFA ⫽ 0.75)
f ⫽ 28 (600 cpsi, OFA ⫽ 0.79)
Pseudo-homogeneous model:
f TP ReL ⫽ const

Lebens et al. (1999) dp


ⵜ共␮i ⵜUiZ 兲 ⫽ ⫹ g␳i (19)
dz
with i ⫽ G, L

relations reported by researchers working in the field of mono- pressure drop attributed to aeration of liquid slugs, which
lith. depends on the number of bubbles formed in the liquid slugs
The pressure drop through the monolith is primarily caused (Table 2). If we assume the phases moving along the monolith
by factors such as (1) the wall friction, (2) the acceleration of channel to be incompressible, the pressure drop caused by the
gas phase, (3) the orifice effect at the entry region and between acceleration can be neglected.
the monolith stacks, and (4) pressure drop caused by the In all the correlations listed in Table 2, the pressure drop
gas–liquid distributor. Mewes et al. (1999) also considered the attributed to wall friction has been modeled in line with Hage-

AIChE Journal November 2004 Vol. 50, No. 11 2923


n–Poiseulle law. For specific monoliths, the friction factors researchers in this field to predict that the ratio between the
have been calculated experimentally as well as theoretically average velocity of the liquid slugs and the average velocity of
(assuming laminar flow) with good agreement. In general, the bubbles differs only slightly from one. Thus, the average vol-
pressure drop attributed to the gas phase was found to be very ume fraction of gas equals approximately the ratio of the flow
small. rate of gas to total flow rate, with a difference of about 15%. It
Considerably less attention has been given to the last two is therefore obvious that the proper measurement of the length
factors (orifice effects and distributor pressure drop). Satter- of the liquid and gas slug would give a fairly reliable measure
field and Özel (1977) used the Bernoulli relationship to model of the holdup.
pressure drop resulting from orifice effects. However, the au- Ishii (1977) investigated two-phase flow in round tubes and
thors found this contribution to be small compared to that of developed the drift-flux (D-F) model to describe the relative
friction losses. motion between phases in flow regimes. According to the D-F
Heiszwolf et al. (2001a) explained the phenomenon of bub- model, the relationship between the gas velocity and the mix-
ble formation in the liquid slugs and its effect on the pressure ture’s volumetric flux is expressed in Eq. 22 (Table 3). Later,
drop. When gas bubbles in the channel are entirely contained Mishima et al. (1996) provided a new distribution parameter
within the liquid and do not interact with the wall, the pressure for two-phase flow in small diameter vertical tubes. This pa-
drop is high because of higher friction of liquid. When enough rameter was obtained from experiments in tubes with inner
gas is introduced, two counteracting effects take place. The diameters ranging from 1 to 4 mm, operated in bubbly flow and
friction factor decreases because of gas friction with the wall slug-flow regimes.
(which is less than liquid friction), but it also, to some extent, Various techniques have been used to measure the lengths of
reduces the laminar nature of the liquid slugs and thus increases gas and liquid slugs. Hatziantoniou et al. (1984) and Irandoust
the friction factor. The authors proposed a two-phase friction et al. (1988a) used two different techniques in a cocurrent
factor fTP to model the pressure drop. downflow mode. First, they measured the displacement of the
Clearly, there is a greater need to investigate the entrance pump piston, which gives the liquid slug length, and measured
and exit effects, particularly the role of gas bubbles of varying the frequency of the stroke for the gas bubble length. Second,
size on the pressure drop. Researchers have used various types they used of a conductivity cell, whose response is shown in
of distributors in conducting experiments related to monoliths. Figure 4. Two probes were installed at the exit of a monolith
Most often, however, only the pressure drop across the mono- channel. One was placed at the center of the channel, the other
lith has been reported, which could lead to undersizing of one at the channel wall.
pumps and compressors while designing a monolith unit. The results obtained by the two methods were within 9.5%
Finally, it is important to investigate pressure drop in flow of each other. It was also observed that the lengths of the gas
regimes other than slug flow. In some of the recent investiga- and liquid slugs could be independently varied by changing the
tion related to monolith, researchers are investigating the ap- corresponding flow rates.
plication of monoliths in annular flow (Roy et al., 2002). In subsequent works, Irandoust et al. (1989, 1992) used a
photocell to visualize the flow regime and to measure the plug
lengths in a glass capillary of diameter 2.2 mm. The lengths of
Holdup the gas slugs varied from 3.4 to 29.1 mm, and those of the
Gas–liquid holdup is an important hydrodynamic parameter liquid varied from 2.9 to 67 mm, for a total average linear
for reactor design, scale-up, and performance modeling. Con- velocity of 0.092 to 0.56 m/s.
siderable research has been performed on packed-bed and Grolman et al. (1996) performed direct integral holdup mea-
trickle-bed reactors to determine phase holdup, and reviews are surements. The liquid holdup was obtained by continuously
available in the literature (see, for example, Al-Dahhan et al., weighing the column. Furthermore, a model was presented that
1999). Several methods have been used to measure the holdup, takes into account the liquid slug and the liquid film between
including gravimetric method (Nemec et al., 2001) for overall the gas bubble and the wall. The experimental values are in
holdup and the tomographic method for cross-sectional holdup good agreement with the model prediction for liquid holdup
(Boyer et al., 2002; Chen et al., 2001). Determination of holdup values higher than 65%. The authors assume that the discrep-
and holdup distribution is as important in structured beds as it ancies at lower liquid holdup are caused by liquid maldistri-
is in packed beds. The flow regimes significantly affect the bution.
holdup. In structures such as sandwich beds and trickle beds, Few researchers have considered holdup measurements in
the regimes are considerably different from those in monolith flow regimes other than Taylor flow. The focus of their inves-
because of the difference in flow paths. In monoliths, the tigations has been the annular or film flow regime in counter-
holdup is characterized by the formation of liquid slugs in the current and cocurrent downflow mode. The annular flow re-
channels in a regular fashion, whereas the liquid holdup in gime is akin to film flow: the liquid trickles down the corner of
trickle beds and other structured bed reactors arises from riv- the channels and the gas occupies the core.
ulets between the particles and structures. Heibel et al. (2001a) successfully used magnetic resonance
In monoliths operating in the Taylor flow (slug flow) regime, imaging (MRI) to measure the holdup inside the monolith in
the liquid holdup is simply the ratio of average liquid slug the film-flow regime. They used a cell density of 25 cpsi and
length to the total length of the liquid and gas slug combined, operated the monolithic reactor in downflow mode. Water
neglecting the contribution of liquid film surrounding the gas velocity was varied from 0.45 to 4 cm/s, and no gas flow was
bubble. This is true when the slugs do not coalesce, which is used. Within the range of flow rates the liquid holdup varied
the case reported by most investigators (Cybulski et al., 1998; between 4 and 15%. The system was also modeled by solving
Grolman et al., 1996). Mewes et al. (1999) quoted previous Navier–Stokes equations, assuming no slip between the gas

2924 November 2004 Vol. 50, No. 11 AIChE Journal


Table 3. Holdup Correlations for Capillaries and Monolith Channels
Author (Year) Holdup Correlation

Wallis (1969) ␾G uG
␧G ⫽ ⫽ ⫽␤ (20)
共␾G ⫹ ␾L 兲 共uG ⫹ uL 兲

Butterworth (1975) 1 ⫺ ␧G
␧G
⫽A 冉 冊冉 冊冉 冊
1 ⫺ x p ␳G q ␮L r
x ␳L ␮G
(21)

For Lockart–Martinelli (1949) correlation:


A ⫽ 1.0, p ⫽ 0.64, q ⫽ 0.36, and r ⫽ 0.07
For Broncy (1963) correlation:
A ⫽ 1.0, p ⫽ 0.74, q ⫽ 0.65, and r ⫽ 0.13
Ishii (1977) v G ⫽ u G /␧ G ⫽ C 0 (u G ⫹ u L ) ⫹ V G,u (22)
uG
␧G ⫽ (23)
C0 共uG ⫹ uL 兲 ⫹ VG,u

Slug flow:
C 0 ⫽ 1.2 ⫺ 0.2 公␳ G / ␳ L (24)
⫺0.691d c
C 0 ⫽ 1.2 ⫹ 0.51e (Mishima et al., 1996) (25)
V G,u ⫽ 0.35 公⌬ ␳ gd c / ␳ L (26)

Grolman et al. (1996) uL ⫹ uG 共1 ⫺ R /R 兲 ⫺ ␾LF /␲R


2
0
2 2

␧L ⫽ (27)
共uL ⫹ uG 兲 ⫺ ␾LF /␲R2

and liquid interface, and the model’s predicted holdup was in also depends on the total pressure of the system, by changing
very good agreement with the MRI results except at low liquid this pressure, the gas velocity and thus the gas holdup can be
flow rates (Ul ⬍ 0.75 cm/s). changed to some extent.
In general, the gas and liquid velocities inside the monolith Unfortunately, most of the research performed in this field
channels can be independently varied by changing the gas and does not address the concern of holdup distribution across the
liquid flow rates into the system. Therefore the holdup can be monolith cross section, which can prove to be an important
changed either by changing liquid or gas flow rates. Broekhuis parameter for reactor design. Recently, noninvasive techniques
et al. (2001) and Heiszwolf et al. (2001a) described a monolith such as capacitance, X-ray and ␥-ray tomography, and MRI
loop reactor in which gas is recirculated within the system have been successfully used in conventional reactors as well as
using an ejector driven by liquid flow. The gas recirculation in structured beds (Harter et al., 2001; Heibel et al., 2001b;
velocity depends on the liquid velocity, the two being related Kumar et al., 1997; Mewes et al., 1999; Reinecke et al., 1998)
because of the equilibrium of pressure drop in the monolith and can be extended to study the holdup distribution in mono-
section and recirculation section. Therefore the gas holdup lith packing.
cannot be independently changed and depends entirely on the
liquid velocity. Because the pressure drop in the two sections Flow distribution
Uniform flow distribution in a multiphase reactor is impor-
tant for enhanced productivity and selectivity. It ensures com-
plete use of the catalyst and prevents hot spots in an exothermic
reaction system. For conventional reactors such as packed
beds, slurry columns, and bubble columns, distributions of the
gas and liquid phases have been studied in detail (Marcandelli
et al., 2000), including various techniques used to measure flow
maldistribution. The same measurement techniques have been
extended to a structured bed as well. The importance of gas/
liquid distribution inside monoliths is much more profound
because, unlike random packing or other structured packing,
once the liquid enters the reactor, there is no further redistri-
bution inside the monolith. Therefore the liquid must be dis-
tributed uniformly before it enters the monolith bed.
Figure 4. Conductivity cell response at three different Several ways have been devised by various researchers for
gas/liquid flow rates. liquid distribution in monoliths operating in gas–liquid cocur-
(a) uL ⫽ 16.7 cm3/s, uG ⫽ 15.6 cm3/s; (b) uL ⫽ 16.7 cm3/s, rent downflow mode. A showerhead was the most widely used
uG ⫽ 27.0 cm3/s; (c) uL ⫽ 10.8 cm3/s, uG ⫽ 15.6 cm3/s,
channel cross section 2.0 mm2, bed porosity 59% (from Hat- distribution device (Hatziantoniou et al., 1984, 1986; Irandoust,
ziantoniou et al., 1984). 1989; Nijhuis, 2001; Satterfield, 1977) because of its easy

AIChE Journal November 2004 Vol. 50, No. 11 2925


Table 4. Summary of Distribution Systems Used by Different Researchers in Monolith
Direction
Author/Year of Flow Distributor Velocities Remarks
Satterfield et al. (1977) Downflow Shower head layer of u L ⫽ 0.3 to 6.6 cm/s Layer of monolith discs gave
spheres monolith reproducible pressure drop
discs values, indicating uniform
distribution
u G ⫽ 0–150 cm/s
Irandoust et al. (1989) Downflow Sieve plates u L ⫹ u G ⫽ 0.27 to
0.53 m/s
Crynes et al. (1995) Upflow Glass frit u L ⫽ 0.4 to 3,5 cm3/ Distributor produces froth,
s which then travels up
u G ⫽ 15.8 to 50 Overall monolith activity
cm3/s achieved was same as
intrinsic rate, indicating
uniform distribution
Reinecke et al. (1996) Downflow u L ⫽ 0–0.2 m/s Maldistribution observed
using capacitance
tomography. With
increasing liquid velocity,
distribution improved
u G ⫽ 0–8 m/s
Mewes et al. (1999) Downflow u L ⫽ 0.26–0.39 m/s Same as above
u G ⫽ 0–5 m/s
Broekhuis et al. (2001) Downflow Liquid-driver ejector u L ⫽ up to 20.4 cm/s The ejector action produces
fine gas bubbles in liquid

construction and operation. Table 4 gives a summary of distri- lithic reactor is considered. At low liquid velocities, the pres-
bution systems used by different researchers. sure drop across the monolith is negative (hydrostatic pressure
Irandoust et al. (1989) evaluated the flow distribution in a is more than the frictional pressure drop), and this condition
monolith operated in downflow by using several perforated forced recirculation of trapped gas inside the monoliths. With
sieve plates, with perforations ranging from 5.7 to 27% open increasing liquid velocity, the quality of liquid distribution
area and hole diameter from 0.5 to 2.75 mm. The experiment increases. The effect of gas velocity on the liquid distribution
was performed in a column having a diameter of 25 cm. Water was not demonstrated.
flowed through the distributor, whereas air made a side entry Crynes et al. (1995) developed a novel “monolith froth
just before the monolith. It was observed that the liquid flow reactor” in which a gas–liquid froth was made to flow up under
distribution was dependent on pressure drop across the perfo- pressure. The froth was prepared by passing gas through a glass
rated plate. The liquid was poorly distributed when the pressure frit (145- to 175-␮m pore diameter) over which liquid was
drop across the plate fell below 100 –200 mm of water and flowed. The system was tested for aqueous oxidation of phenol
resulted in a much thicker spray of liquid over the monolith in a ␥-alumina washcoat monolith impregnated with CuO as
channels. According to Cybulski et al. (1998), the liquid spray active metal. The observed reaction rate was close to the
drops should be much smaller than the channel diameter and intrinsic rate, which the authors interpreted as a sign of good
should be sprayed uniformly over the monolith. phase distribution and minimal transport resistances.
Satterfield et al. (1977) investigated three different types of Broekhuis et al. (2001) used a liquid-motive ejector as a
liquid distributors for downflow arrangement and evaluated the gas–liquid distributor. They claim that this ejector is also a very
results for reproducibility. First, a flat distributor head with an good gas–liquid contactor, presaturating liquid before it enters
arrangement of 37 capillaries, varying from about 3 to 7 mm, the reactor. It produces a fine dispersion of gas bubbles in
was used. The distribution was not uniform over the cross liquid, which results in excellent gas–liquid distribution over
section, which resulted in irreproducible pressure drop data. In the cross section of the monolith. The ejector also acts as a gas
addition, a slight rotation of the head caused significant differ- compressor, resulting in higher superficial gas and liquid ve-
ences in pressure drop. The second distributor was a layer of locities compared to those produced by gravity-driven mono-
4-mm-diameter spheres. Some of the spheres plugged the en- lith reactors.
trance of monolith channels, and flooding above the layer It is interesting to note that, in most of the studies mentioned
occurred at high liquid flow rates. This distributor also gave above, the role of gas velocities on the liquid distribution has
unusable data. The authors obtained best distribution (indicated not been studied. In a critical observation of the published
by reproducibility of pressure drop) by using 27 randomly works, the researchers have mainly depended on indirect meth-
aligned stacks of monoliths, each 3.2 mm thick, below the ods of qualitatively estimating the flow maldistribution, such as
showerhead. consistency in pressure drop measurement (Satterfield et al.,
Mewes et al. (1999) studied flow distribution in monoliths 1977) or high reactor productivity (Crynes et al., 1995). This
by capacitance tomography. A spatial resolution of about has resulted in no single parameter being used to quantitatively
5–10% of the diameter of the measurement plane is in general define a maldistribution and compare the performance of these
possible using capacitance tomography (Reinecke et al., 1999). distributor studied.
In this study, only liquid flow distribution in a 120-mm mono- Modern-day noninvasive tools such as computed tomogra-

2926 November 2004 Vol. 50, No. 11 AIChE Journal


Figure 5. Axial dispersion coefficient for bubble-train flow in 2 mm.
(a) Circular capillary as a function of capillary number (Ca ⫽ ␮v/␴): E, experimental; f, theory. (b) Square capillaries: f, experimental data
(from Thulasidas et al., 1999).

phy, electrical and capacitance tomography, and MRI can help imental work to determine the liquid-phase RTD in capillary
generate quantitative images of the gas and liquid phase on a tubes of circular and square cross section, operated in upflow
cross section and longitudinal sections. These images will lead mode. A corresponding mathematical model was developed,
us to better understand the dynamics of each type of distributor. based on the mass balance of the tracer element. In their model,
complete mixing within the liquid slugs and transport of tracer
Axial and radial dispersion through the liquid film surrounding the gas bubble were as-
sumed. The model and the experimental data were in agree-
Neglecting proper characterization of the mixing phenomena
ment for a circular capillary. It was found that the measured
can result in considerable error in the reactor performance. In
packed-bed reactors, the idealized assumption of plug flow for mean residence time for the circular capillary was 20.0 s,
modeling purposes is no longer pursued and considerable work whereas the time predicted by the model was 19.6 s. The result
has been done to develop various levels of models such as the confirms the small amount of backmixing in circular capillar-
axial dispersion model (ADM), tanks in series model, and so ies. However, the model failed to predict the mean residence
forth (Ramachandran et al., 1983; Sundreasan, 1986; Taylor, time for square channels as well as it did for the monolith froth
1953). These models attempt to account for the flow deviation reactor. The experimentally obtained mean residence time for a
from plug flow character and to improve reactor performance. square capillary was 27.9 s. The authors argue that the liquid
In modeling monolith reactor performance, plug flow of the slugs are not well mixed because the liquid flows through the
liquid slugs has most often been assumed (Cybulski et al., corners of the channels. The liquid flowing through the corners
1993, 1999; Edvinsson et al., 1994; Hatziantoniou et al., 1984; of the channels will almost completely bypass the liquid slugs.
Irandoust et al., 1989b,c; Stankiewicz, 2001). This assumption It is claimed that the model can be extended to the monolith
is bolstered by the fact that in slug flow, the liquid slugs are reactor by applying a statistical analysis of the flow within the
intensely mixed (Irandoust et al., 1989b) and there is very little channels of the monolith.
concentration gradient in either the radial or axial direction if Thulasidas et al. (1999) performed tracer studies in capillar-
no chemical reaction takes place. Moreover, the thickness of ies with circular and square cross sections with a hydraulic
the liquid film around the gas bubbles is very small. There is diameter of 2 mm and also studied a capillary bundle. The
thus no interaction between two successive liquid slugs, and bundle consisted of 96 square capillaries with a hydraulic
hence axial dispersion is minimal. diameter of 2 mm. The studies were carried out in upward slug
In a study of residence time distribution (RTD) in monolith flow mode. A mass transfer model was used to predict con-
froth reactor in cocurrent upflow, Patrick et al. (1995) observed centration vs. time curves for liquid slugs leaving the capillar-
that interaction takes place as a result of dispersion in the thin ies. The results show good agreement between the model
liquid film surrounding the gas slugs. A series of tracer studies, prediction and the experimental data. Axial dispersion coeffi-
by use of 4,6-dichlororesorcinol (DCR), was conducted in a cients computed from experimental values of the Peclet num-
monolith reactor assembly with a cell density of 400 cpsi. The bers (Pe) for bubble train flow in circular and square capillaries
actual liquid residence time in the monolith was calculated are shown in Figures 5a and 5b.
from overall residence time measurements, using deconvolu- The theoretical model was extended for a bundle of capil-
tion by Fourier transform. When fitted into the experimental laries and used to estimate residence time distributions. Nor-
RTD data obtained at a liquid flow rate of 4.7 cm3/s and a gas malized concentration vs. time curves for the capillary bundle
flow rate of 47 cm3/s, a tank-in-series model predicted the are shown in Figure 6a; the computed RTD for a capillary
monolith as 1.15 perfectly mixed vessel, which indicates a high section is shown in Figure 6b. The average residence time from
degree of backmixing inside the monolith. the distribution shown in the latter figure is 69.41 s. The model
Thulasidas et al. (1995b) also reported modeling and exper- predicts well the RTD in a single capillary but for an extension

AIChE Journal November 2004 Vol. 50, No. 11 2927


Figure 6. (a) Experimental normalized concentration distribution for liquid-only flow and bubble flow in a capillary
bundle: f, liquid flow only; E, bubble train flow; (b) normalized concentration vs. time distribution for
capillary bundle after deconvolution; f, experimental tracer concentration vs. time distribution (from
Thulasidas et al., 1999).

to capillary bundles or monoliths it needs further studies of the gas–liquid mass transfer coefficient (kGL). The transport of
flow inside these arrays. liquid reactant and the dissolved gas in the liquid from the bulk
Obviously, there is an apparent disagreement among re- of the liquid to the washcoat surface depends on the liquid–
searchers about the presence of axial dispersion in monoliths. solid mass transfer coefficient (kLS). Gas in the gas slug also
Even though cocurrent downflow operation has been assumed diffuses to the surface of the catalyst through the thin liquid
to be plug flow, direct experimental verification of this assump- film, which is characterized by the gas–solid mass transfer
tion is not available. In almost all studies, the regime of interest coefficient (kGS). Finally, the reactants on the surface of the
was slug flow (Taylor flow). However, investigating peripheral catalyst have to diffuse into the pores of the washcoat or the
regions such as bubbly flow and churn flow will be of interest catalyst wall by standard diffusion processes. This diffusion
for many mass transfer limited reactions. For annular flow, the has been dealt with extensively in standard reaction engineer-
flowing film is considered laminar and treated accordingly ing text books (Fogler, 1992; Froment and Bishoff, 1990;
during modeling (Roy et al., 2002). Levenspiel, 1998) and will not be discussed in detail here. All
the above information is needed to adequately model the over-
Mass Transfer all performance of the reactor. In subsequent sections, we will
explore each of the transport properties in detail.
As mentioned earlier, one of the most attractive features of
monolith as compared to random packing for a multiphase
reaction is its enhanced mass transfer characteristics. In the Liquid–solid mass transfer
Taylor flow regime, an intense mixing takes place within the Hatziantoniou et al. (1982) studied the liquid–solid (L-S)
liquid slug, which significantly enhances the liquid–solid mass mass transfer in capillary tubes of 2.35 and 3.094 mm diame-
transfer rate. Moreover, the film separating gas plugs and ters. The channel wall was coated with benzoic acid, and water
channel walls is very thin, posing minimal gas–solid mass and air were passed cocurrently downward. The outlet concen-
transfer resistance. Most of the correlations for liquid–solid tration of benzoic acid dissolved in water was measured. The
mass transfer developed so far in random packing are based on L-S mass transfer coefficient was extracted from experimental
power-law correlations relating the Sherwood number with the results and by performing a differential mass balance on ben-
Reynolds number, Schmidt number, and other geometrical zoic acid assuming a plug-flow behavior. Based on the exper-
properties of mass transfer devices (Highfill et al., 2001). imental results, the author proposed the following correlation
Because the mechanism of mass transfer is expected to be the for the liquid–solid mass transfer coefficient
same in monoliths, the same dimensionless groups may play
important roles in the development of mass transfer correla-
tions in monoliths.
The overall mass transfer rate in monolith depends primarily
on four different mass transfer phenomena, as depicted in
Figure 7. Because the reaction takes place on the surface and
inside the pores of the washcoated wall, the gas and liquid Figure 7. Different transport phenomena in Taylor flow
reactants have to transfer from the bulk to the catalyst surface. inside a monolith.
The gaseous reactant diffuses from the bulk of the gas slug to Gas/solid (G-S), gas/liquid (G-L), liquid/solid (L-S), pore
the liquid bulk, and the corresponding flux depends on the diffusion.

2928 November 2004 Vol. 50, No. 11 AIChE Journal


Sh ⫽ 3.51 冉 冊
ReSc

0.44
␤⫺0.09 (28)
As is evident, most of the correlations are dependent on the
liquid slug length. However, there is no method available to
estimate the slug length a priori. The individual researchers
have based their experimental results on their own observations
where ␥ ⫽ Lc/dc and ␤ ⫽ Ls/dc.
of slug lengths, and may differ from one study to another.
It is interesting to note that the expression contains the tube
Most of the experiments were conducted on a smooth sur-
length, which makes it unsuitable for scale-up. The flow rate at
face and in clearly defined slugs of liquid. However, in an
which the expression was developed was significantly high so
actual reactor, the liquid slugs may be aerated, giving rise to a
as to maintain complete recirculation within the liquid plugs,
“disturbance” on the liquid–solid interface. Moreover, mono-
and thus the correlation cannot be used at low Reynolds num-
liths likely to be used in industrial settings will seldom have
bers.
smooth surfaces. These factors may cause deviations from
Based on a dissolution study in cylindrical capillaries, Iran-
model predictions (Nijhuis et al., 2001). Table 5 summarizes
doust and Andersson (1988) developed a new correlation to
reported correlations for liquid–solid mass transfer coefficients.
estimate the liquid–solid mass transfer coefficient

Sh ⫽ 1.5 ⫻ 10⫺7 共Re兲1.648 共Sc兲0.177 共␣兲⫺2.338 (29) Gas–liquid mass transfer


Gas–liquid mass transfer in nonstructured trickle-bed reac-
where ␣ is the dimensionless film thickness, ␦f/dc. The authors tors has been studied in great detail by past researchers. Cor-
cited data from an unpublished work to formulate the above relations that cover wide ranges of Reynolds number (Re) and
correlation. Schmidt (Sc) numbers are available in the literature (for a
Bercić et al. (1997) carried out an experiment similar to that review, see Ramachandran et al., 1983). Although the mecha-
of Hatziantoniou et al. (1982), in a glass tube of 2.5 mm nism of transport is not expected to be significantly different in
diameter coated with benzoic acid. Air and water were passed structured packing, the work in this field is somewhat limited
through the tube in a controlled manner to form Taylor flow and incoherent (Heiszwolf et al., 2001b).
inside the tube. Nonlinear regression was used to fit the exper- Similar to their liquid–solid mass transfer work, Bercić et al.
imental data to a model equation to correlate the liquid–solid (1997) proposed a dimensional correlation for the gas–liquid
mass transfer coefficient with the unit cell length (UCL, the mass transfer coefficient. Gas–liquid mass transfer coefficients
sum of one gas and one liquid slug length) and the cell velocity v were measured by physical absorption of methane in water,
using capillary tubes of three different diameters (1.5, 2.5, and
0.069v 0.63 3.1 mm). Methane and water formed the gas and liquid slugs,
k LSa ⫽ (30)
关共1 ⫺ ␧ G兲UCL ⫺ 0.105UCL␧ G兴 0.44 respectively. The measured data were fitted into the proposed
expression, which was developed by assuming plug flow of the
Based on mass transfer experiments, Heiszwolf et al. liquid slug and then using nonlinear regression as follows
(2001b) proposed a semiempirical liquid–solid mass transfer
model 0.111v 1.19
k GLa ⫽ (33)
关共1 ⫺ ␧ G兲UCL兴 0.57

Sh ⫽ 3.66 1 ⫹ 0.152
⌿s
ReSc 冉 冊 册 ⫺0.423
(31)
Because the expression does not contain any physical pa-
rameters of the gas and liquid used, it lacks generality and can
where ⌿s is the dimensionless liquid slug length LS/d. Unfor-
be used only for methane–water systems. Heiszwolf et al.
tunately, the basis for such a correlation remains unpublished.
(1999) attempted to make the expression compatible with other
However, the authors compared the above correlation with that
systems by adding the following correction factor
proposed by previous researchers (such as Eqs. 28, 30, and 31).
The comparison shows that all the correlations more or less
exhibit the same trend and are in the same range. However, the
values predicted by Heiszwolf are slightly higher than those
predicted by other researchers. The authors noted that the
共k GLa兲 A ⫽ 共k GLa兲 B 冉 冊
DA
DB
n
(34)

values predicted by Eq. 29 (Irandoust and Andersson, 1988a)


were not in the range predicted by other correlations. where n is the scaling factor: n ⫽ 1 for film theory, and n ⫽ 0.5
In a recent article Kreutzer et al. (2001) reported heat trans- for penetration theory (Kreutzer et al., 2001). However, this
fer studies using CFD simulation in a 1-mm-diameter tube and modification has its own drawbacks and could not achieve the
developed a correlation relating the Nusselt number (Nu) as a desired results (Heiszwolf et al., 1999).
function of the dimensionless liquid slug length (⌿s), Reynolds In earlier work, Irandoust and Andersson (1988a) used the
number (Re), and Prandtl number (Pr) within the range 1 ⬍ ⌿s following correlation (Eq. 35) to compute kGL as a needed
⬍ 16, 7 ⬍ Pr ⬍ 700, and 10 ⬍ Re ⬍ 400. By analogy, the parameter for modeling the overall performance of a monolith
authors proposed the following correlation for the liquid–solid reactor. The authors cited unpublished work to support their
mass transfer coefficient claim


Sh ⫽ 20 1 ⫹ 0.003 冉 冊 册
⌿s
ReSc
⫺0.7
(32)
Sh ⫽ 0.41 冑ReSc (35)

AIChE Journal November 2004 Vol. 50, No. 11 2929


Table 5. Available Correlations for Liquid–Solid (L-S) and Gas–Liquid (G-L) Mass Transfer
Correlation
Author (Year) Liquid–Solid Mass Tranfer Gas–Liquid Mass Transfer

Hatziantoniou et al. (1982) Sh ⫽ 3.51 冉 冊


ReSc

0.44
␤⫺0.09
with ␥ ⫽ L c /d c and ␤ ⫽ L s /d c
Irandoust et al. (1988a) Sh ⫽ 1.5 ⫻ 10⫺7(Re)1.648(Sc)0.177(␣)⫺2.338 Sh ⫽ 0.41公ReSc
with ␣ ⫽ ␦ f /d c
共Sh ⫺ 1兲
冋 冉 冊册
1/3
Irandoust et al. (1992) 1
1 ⱕ Re ⱕ 400: ⫽ 1⫹ Re0.41
Sc1/3 ReSc
100 ⱕ Re ⱕ 2000: Sh ⫽ 1 ⫹ 0.724 Re0.48Sc1/3
0.069v0.63 0.111v1.19
Bercić et al. (1997) kLS a ⫽ kGL a ⫽
关共1 ⫺ ␧G 兲UCL ⫺ 0.105UCL␧G 兴0.44 关共1 ⫺ ␧G 兲UCL兴0.57

Heiszwolf et al. (1999) 冋


Sh ⫽ 3.66 1 ⫹ 0.152 冉 冊 册
⌿s
ReSc
⫺0.423
DA n
共kGL a兲A ⫽ 共kGL a兲B
DB 冉 冊
n ⫽ 1.0 for film theory
n ⫽ 0.5 for penetration theory (Kreutzer et al., 2001)
0.093␨⫺0.87
Lebens et al. (1999a) Sh ⫽ 1.04 ⫹
1 ⫹ 0.047␨⫺0.61
Lc DL
with ␨ ⫽ 2
dfM uLM
Kreutzer et al. (2001)

Sh ⫽ 20 1 ⫹ 0.003 冉 冊 册
⌿s
ReSc
⫺0.7

In a more recent paper, Irandoust et al. (1992) assumed the determined by measuring desorption by nitrogen gas flow of
hemispherical caps of the gas bubble to be rigid spheres and oxygen from oxygen saturated water. The authors also devel-
cited previous literature on mass transfer correlations for rigid oped a correlation by solving the convection– diffusion equa-
spheres to propose the following gas–liquid mass transfer co- tion for a smooth laminar falling liquid film flow on a vertical
efficient plane, as follows

共Sh ⫺ 1兲
冋 冉 冊册
1/3
1 0.093␨⫺0.87
1 ⱕ Re ⱕ 400: ⫽ 1⫹ Re0.41 (36) Sh ⫽ 1.04 ⫹ (38)
Sc1/3 ReSc 1 ⫹ 0.047␨⫺0.61

100 ⱕ Re ⱕ 2000: Sh ⫽ 1 ⫹ 0.724 Re0.48 Sc1/3 (37) with

Irandoust et al. (1992) performed the experiments in a single L cD L


capillary tube with three different liquids: water, ethanol, and ␨⫽ (39)
ethylene glycol. Air and the liquid were passed cocurrently ␦ f,m
2
u L,m
upward, and the outlet concentration of oxygen in the liquid
was measured for model validation. However, the model over- where Lc is the tube length, DL is the diffusivity, ␦f,m is the
estimated the experimental results by about 30% and therefore maximum film thickness, and uL,m is the maximum liquid
Eq. 37 should be multiplied by a factor of 0.686 before it is velocity.
used elsewhere. It must be noted that the Sherwood number is a function of
During the last decade, processes with two-phase cocurrent the dimensionless tube length (␨), and for higher values of ␨ it
upward or downward flow have been well investigated in approaches the asymptotic value of 1.04. The experimental
monolith reactors. Different researchers have provided a vari- values in general had good agreement with the above correla-
ety of experimental and theoretical data. Recently a new type tion (Eq. 38), except at the inlet region of the tube, where the
of structure, the internal finned reactor, was presented by correlation underestimates the experimental values by a wide
Lebens et al. (1997). This structure enables countercurrent margin.
operation in the monolith. The monolith reactor operates in the The area of contact for gas–liquid mass transfer has always
annular flow regime, and the fins help to stabilize the liquid been evaluated by assuming hemispherical plugs. However, in
film flow. The mechanism of mass transfer is therefore signif- reality, the shape is more complex and varies with capillary
icantly different from monoliths operated in the slug-flow number (Cybulski and Moulijn, 1998).
regime. Despite all the work performed to develop mass transfer
Lebens et al. (1999a) reported both experimental and pre- correlations, they have yet to be used effectively in models of
dicted results on data for gas–liquid mass transfer coefficients. monolith reactor performance at pilot plant or industrial scales.
In the experiment, the gas–liquid mass transfer coefficient was Nijhuis et al. (2001) used the mass transfer correlation devel-

2930 November 2004 Vol. 50, No. 11 AIChE Journal


oped by Irandoust et al. (1992) to model a pilot scale reactor. favor fast reactions, studies of intrinsic slow reactions also help
However, the reactor performance predicted by the model far gauge the effectiveness of the monolith.
overestimated the actual performance in experiments. The au- Most of the researchers, while evaluating the performance of
thors argue that in actual reactors the resistance layer is much monolith, have compared it with that of conventional reactors.
thicker than that in model assumptions. Moreover, the wash- There has been a general disagreement about the basis on
coated channel walls are irregular, which would contribute to which such comparison should be made, as discussed in the
the increase in mass transfer resistance. section on performance comparison. Finally, the models pro-
As mentioned in the previous section, aeration of the slug posed by investigators are outlined in the section on reactor
may introduce error in operation near the churn flow regime. A modeling. These models are based on the mass balance of the
comprehensive study is needed to account for all of these reacting species, taking into account different mass transport
factors, so that the correlations can be used in overall modeling phenomena occurring inside the packing.
with some degree of confidence. Table 5, found at the end of
this section, lists the available correlations for liquid–solid and Test chemical reactions used
gas–liquid mass transfer coefficients.
Monolith reactors have primarily been used to catalyze au-
tomobile engine exhaust. More recently, multiphase reactions
Gas–solid mass transfer such as hydrogenation, and partial and complete oxidation
Researchers have evaluated the gas–solid mass transfer in a reactions, which are traditionally performed in slurry, fluidized,
simplistic manner. The gaseous reactant diffuses to the solid or packed-bed reactors, are being tried in structured packing,
surface through the thin film surrounding the gas slug. The such as monoliths, sandwich or open cross-flow structures, and
thin-film model was therefore used to determine the gas–solid foam structures.
mass transfer coefficient (Cybulski et al., 1993; Edvinsson et The advantages of structured packing lie in the field of fast
al., 1994; Irandoust et al., 1988) reactions, which are mass transfer limited. The intense mixing
in the segmented slug flow, together with the very thin liquid
D film between the gas bubbles and the catalyst surface, leads to
k G-S ⫽ (40) enhanced mass transfer. In studying the overall performance of
␦f
monolith reactors, the researchers have chosen a particular
reaction for a variety of reasons. Some reactions represent
Therefore, to determine the gas–solid mass transfer coefficient, industrially relevant processes or exemplify a group of similar
one has to evaluate the film thickness (␦f) of the liquid film, reactions; other reactions have well-known kinetics and sub-
which depends primarily on the surface tension of the liquid stantial amounts of experimental and theoretical data available.
and the liquid diffusivity (D). However, the film thickness is Most of the research efforts has centered on hydrogenation
not uniform around the periphery of the monolith channel, and reactions, one of the most important reactions in the petroleum
a mechanism to account for this nonuniformity needs to be and fine chemical industries. These reactions are typically a
designed. It must be noted that experiments to directly deter- three-phase reaction over a catalyst surface, in which hydrogen
mine the gas–solid mass transfer coefficient are not available in reacts with liquid in the presence of a solvent, which improves
the literature. heat removal. Another type of reaction, catalytic wet oxidation,
can be used to purify industrial and human wastewater. Such
Reactor Performance processes are characterized by the reaction of oxidizable liquid
or solid with a gaseous source of oxygen, such as air or pure
So far we have dealt with the hydrodynamics and the trans-
oxygen, to produce CO2 and other innocuous end products
ports inside monolith packing being operated in a range of
(Mishra, 1995). Table 6 gives a summary of research per-
operating conditions. These studies help us to understand the
formed using test reactions in monolith beds. As the table
basic characteristics of monolith packing under cold flow con-
suggests, most reactions studied in monolith reactors are either
ditions—with no chemical reactions. Many unit operations,
hydrogenation or oxidation reactions, which are normally in-
such as absorption, stripping, and distillation, can be modeled
trinsically fast reactions. This is the type of reactions monolith
better with these basic understandings.
is most suited for, given that enhancing mass transfer is what it
The primary objective of this work is to review the effec-
does best. It cannot enhance the intrinsic rate. On certain
tiveness of monoliths for gas–liquid–solid reactions. Currently,
occasions, very slow reactions have also been studied, to com-
the conventional reactors, such as slurry, fluidized-bed, and
pare the intrinsic reaction rate and overall reaction rate in a
packed-bed reactors have been used in industry to carry out
monolith reactor.
three-phase reactions. For a decade now, researchers have
shown that structured packing could be a viable alternative to
that for mass transfer limited reactions. However, to effectively Performance comparison between monolith and
compete with conventional reactors, a wide range of reactions conventional reactors
need to be studied in structured packing, and an adequate The focus of most investigations of monolith and other
model to predict performance also needs to be devised. structured packing for three-phase reactors is to evaluate their
This section covers work on these topics by various re- performance vs. traditional reactors such as slurry bubble col-
searchers. It includes investigations of different test chemical umns, trickle-bed reactors, and batch reactors. Comparison has
reactions (see the following subsection). The reactions were been made both experimentally and by using model predic-
chosen for their industrial relevance or for being intrinsically tions. Table 7 provides a summary of the comparison studies.
fast or slow. Although it is well known that monolith structures Frequently, the basis of comparison is the extent of conversion

AIChE Journal November 2004 Vol. 50, No. 11 2931


Table 6. Summary of Reported Investigations Performed Using Test Reactions in Monolith Beds
Author Reaction Catalyst/Support Conditions Velocities/Mode of Operation
Hatziantoniou et al. (1984) Hydrogenation of nitrobenzoic acid 2.5 wt % Pd 310, 353 K u tot ⫽ 3.03–5.1 cm/s
1, 4 bar downflow
Hatziantoniou (1986) Hydrogenation of nitrobenzene and 5.4 wt % Pd 246, 376 K u tot ⫽ 1.7–4.2 cm/s
m-nitrotoluene SiO2 5.9, 9.8 bar downflow
Mazzaroni (1987) Hydrogenation of ␣-methylstyrene 1 wt % Pd 303–323 K u L ⫽ 0.05–0.34 cm/s
0.2–1 bar u G up to 0.12 cm/s upflow,
downflow
Irandoust (1988a) Hydrogenation of 2-ethylhexenal 0.5 wt % Pd 413, 433 K u tot ⫽ 2.3–8.5 cm/s
SiO2 4.0, 9.8 bar downflow
Kawakami (1989) Oxidation of glucose Glucose oxidase 298 K u L ⫽ 0.004–0.25 cm/s
u G ⫽ 0.5–6.0 cm/s upflow,
downflow
Irandoust (1990) Hydrodesulfurization of thiophene 12 wt % Co 509, 523 K u tot ⫽ 1.72–2.13 cm/s
and hydrogenation of 4 wt % Mo 30, 40 bar downflow
cyclohexene ␥-Al2O3
Edvinsson (1993) Hydrodesulfurization of C–Mo 543, 558, 573 K u tot ⫽ 4.6 cm/s downflow
dibenzothiophene ␥-Al2O3 60, 70, 80 bar
Edvinsson (1995) Hydrogenation of acetylene 0.04 wt % Pd 303, 313 K u L ⫽ 0.66 cm/s
␣-Al2O3 13, 20 bar u G ⫽ 6.5 cm/s downflow
␥-Al2O3
Crynes (1995) Oxidation of aqueous phenol CuO 383–423 K u L ⫽ 0.4–3.5 cm/s
␥-Al2O3 4.8–11.7 bar u G ⫽ 15.8–50 cm/s upflow
Smits (1996) Hydrogenation of a mixture of Pd 316, 343 K u tot ⫽ 5.0–45.0 cm/s
styrene and 1-octane in toluene ␣-Al2O3 5–15 bar downflow
Klinghoffer (1998a,b) Wet oxidation of acetic acid Pt 385–573 K u L ⫽ 0.024–0.093 cm/s
Al2O3 3.5–128 hour u G ⫽ 2.36 cm/s (SCM)
upflow
Patrick (2000) Wet oxidation of glucose and 0.26 wt % Pt 333–455 K u L ⫽ 0.024 cm/s
cellulose Al2O3 u G ⫽ 1.48 cm/s (SCM)
upflow
Nujhuis (2001) Hydrogenation of ␣-methylstyrene 1 wt % Ni 373, 423 K u tot ⫽ 20 cm/s downflow
hydrogenation of benzaldehyde 9 wt % ␥-Al2O3 10, 15 bar
Schutt (2002) Wet oxidation of cellulose 0.34 wt % Pd 373–428 K u L ⫽ 0.026 cm/s
Al2O3 u G ⫽ 2.2 cm/s (SCM)
upflow
Liu (2002) Dehydrogenation of ethylbenzene 72 wt % 866, 878 K u L ⫽ 0.0014 cm/s downflow
␣-Fe2O3 1 bar
16 wt % K2O
4 wt % CeO2

under similar operating conditions, productivity, selectivity, ever, the mass of catalyst, mass of active metal, and external
and pressure drop. The volume of catalyst was the primary surface area have also been considered in the comparisons.
weighting factor for comparison of different reactors. How- Compared to the conventional reactors, structured packing

Table 7. Summary of Studies on the Comparisons between Monolith and Conventional Reactors
Author Comparison Kind of Study Reaction
Mazzaroni et al. (1987) Monolithic reactor vs. trickle-bed Experimnental studies Hydrogenation of ␣-
reactor methylstyrene
Cybulski et al. (1993) Monolithic reactor vs. literature Mathematical model of Liquid-phase methanol
data of slurry columns, monolith reactor synthesis
autoclaves, and trickle-bed
reactors
Edvinsson et al. (1994) Monolithic reactor vs. trickle bed Numerical simulations Three-phase hydrogenation
Cybulski et al. (1999) Monolithic reactor vs. agitated- Mathematical modeling of Hydrogenation of 3-
slurry reactor both reactors hydroxypropanal
Nijhuis et al. (2001) Monolithic reactor vs. trickle-bed Pilot-scale study Hydrogenation of ␣-
reactor methylstyrene and
benzaldehyde
Stankiewicz et al. (2001) In-line monolith reactor vs. Theoretical studies
trickle-bed reactor
Heiszwolf et al. (2001a) Monolith loop reactor vs. bubble Theoretical studies
column and slurry reactor
Roy (2002) Generic method to design Theoretical studies
monolith catalyst
Monolith reactor vs. pellet-based
trickle-bed reactor

2932 November 2004 Vol. 50, No. 11 AIChE Journal


leads to differences in catalyst loading, mass transfer resis-
tance, contacting area, and pressure drop because of the regu-
larity. There is no doubt that the new structures have a superior
advantage with respect to pressure drop. The other parameters,
however, have to be investigated further and optimized to
guarantee a better overall performance than that of traditional
reactors. Unfortunately, various investigators have used differ-
ent criteria for the comparison of monoliths with traditional
reactors.
Mazzarino and Baldi (1987) investigated hydrogenation of
␣-methylstyrene into cumene using a ceramic monolith coated
with Pd. Comparison was made with a trickle-bed reactor
(TBR) based on catalyst weight, and the monolith reactor was
found to perform better than the TBR at high gas flow rates,
whereas the reaction rate was virtually insensitive to the liquid
flow rate. The monolith performed better in upflow mode than
in downflow mode because of better wetting of the catalyst. Figure 9. Attainable selectivity and production for
Figure 8 shows the hydrogenation rates per unit Pd mass of MASR, MR, and TBR (from Cybulsky et al.,
␣-methylstyrene in the monolith reactor and the TBR at dif- 1999).
ferent gas flow rates under the same operating conditions. It
should be noted that better mass transfer in monolith is attrib- Edvinsson et al. (1994) used numerical simulation to quan-
uted to recirculation within the liquid slugs and to the thin titatively compare the performance of a conventional trickle-
liquid film separating the gas slug from the catalyst layer, as bed reactor and a monolith reactor, based on the catalyst load.
mentioned earlier. To demonstrate this point, the authors con- The monolith reactor had washcoated catalyst with varying
ducted a single-phase reaction, in which the liquid was satu- washcoat thickness (resulting in catalyst fraction between 6 and
rated with hydrogen before entering the reactor. In this circum- 33%), whereas catalyst particles used in the simulated TBR had
stance the TBR performed better than monolith in terms of eggshell with a thickness corresponding to approximately the
overall conversion. same catalyst loading (tshell values of 0.5 to 0.05dp). The model
Cybulski et al. (1993) developed a mathematical model for reaction was a consecutive three-phase hydrogenation. For the
liquid-phase methanol synthesis. Different process conditions model system, the selectivity of the intermediate product was
and design parameters were simulated. It was found that the found to be higher for the monolith over almost the whole
optimum thickness of the catalyst layer is in the range of 50 –75 range of variables studied. For particles larger than 2 mm, the
␮m. The authors found that the performance of a monolith volumetric space–time yield was higher in the monolith reactor
reactor is not always superior to that of a conventional reactor. compared to that in the TBR. For particles smaller than 2 mm,
The monolith reactor’s performance was better with respect to however, the pressure drop in the TBR is considerably higher.
CO conversion and methanol productivity per unit volume of The pressure drop in the MR was sufficiently low that gas
the reactor when a stoichiometrically balanced syngas was used could be recirculated without the requirement of external
as feed. The improvement stems from the fact that the monolith pumping.
reactor used contained a theoretically higher concentration of Cybulski et al. (1999) conducted a comparative study of
catalyst than did the slurry reactor. In the case of the TBR, even hydrogenation of 3-hydroxypropanal to 1,3-propanediol in a
though the catalyst concentration is higher, the intraparticle mechanically agitated slurry reactor (MASR) and a monolith
diffusional resistance is significantly higher. reactor. A purely numerical study was done to compare the
performance of the MR and MASR based on productivity
(mol/m3 s⫺1) and selectivity. Although the comparisons were
made on the basis of per unit reactor volumes, the authors
emphasized the point that much more catalyst can be packed in
an MR than in an MASR per unit volume. They adopted a
modeling approach similar to that used by Edvinsson et al.
(1994), discussed earlier. Intrinsic kinetics, mass transfer cor-
relations, and model equations derived earlier by various re-
searchers (Andersson, 1998; Zhu, 1995) were directly used in
this study. As shown in Figure 9, the numerical studies clearly
demonstrated the superior performance of the MR over MASR,
with respect to reactor productivity and process selectivity.
It should be noted that the actual productivity of a MASR is
higher than that shown in the figure. However, given that an
MASR requires considerable downtime for filling, emptying,
and filtering, the overall productivity decreases significantly.
Figure 8. Conversion rates per unit palladium mass of Nijhuis et al. (2001) did a comparative study of hydrogena-
hydrogenation of ␣-methyl styrene vs. gas ve- tion of ␣-methylstyrene and benzaldehyde in a trickle-bed
locity at 313 K (from Mazzarino et al. (1987). reactor and a monolith. Experimental results showed that the

AIChE Journal November 2004 Vol. 50, No. 11 2933


monolith reactor had a 50% higher activity per unit reactor For stirred tanks and bubble columns, previously reported
volume than that of the trickle-bed reactor (Table 8), even correlations for P/V were used (Schluter et al., 1992; van’t Riet,
though the porosity of the monolith reactor is relatively high— 1979). For low P/V, It was found that the kGLa value for the
this despite the fact that the intrinsic rate for the extruded monolith configuration was much higher than that for other
catalyst used in the TBR was slightly more than that of the MR conventional reactors. For high P/V (⬎10,000 W/m3), all types
catalyst. This finding clearly demonstrates the fact that the of reactor give comparable kGLa values. It was further observed
better performance in the MR is entirely attributed to its struc- that for monolith reactors, the kGLa value is a weak function of
ture. Experiments also revealed that the enhanced MR perfor- P/V, meaning the mass transfer coefficient remains almost
mance resulted from the fact that Taylor bubbles in an MR constant for significantly high gas and liquid flow rates.
form a thin film through which most of the mass transfer takes
place. If, instead, the liquid reactant is saturated with hydrogen
and the resultant single phase is passed through the MR, the Reactor modeling
activity falls to about one fourth of the three-phase Taylor flow Because of its well-defined structure, the modeling of mono-
regime. lith is considered more straightforward than the modeling of
MRs are also known to exhibit higher selectivity because of random packed beds. However, the environment outside the
the narrow distribution of their residence time. The hydroge- monolith (distributor, outlet arrangements) affects its perfor-
nation of benzaldehyde showed significantly higher selectivity mance to a great extent. This section deals with different
(⬎90%) with respect to benzylalcohol in an MR, whereas a models reported in the literature to simulate the performance of
TBR gave a selectivity of 73%, thus reaffirming the advantage monolith reactors. In most published work related to monolith,
of the MR over the TBR. maximum emphasis has been given to mass transfer correla-
Stankiewicz (2001) proposed a novel concept in which the tions and pressure drop because these packings have been used
conventional reactor vessel is replaced with monolith units, in industry for over two decades as mass transfer elements,
installed in the process pipelines. The reactor configuration is such as in distillation and absorption.
known as an “in-line monolith reactor” (ILMR). The author Table 9 lists the models for monolith reactors reported in the
made comparison studies for a slow hydrogenation reaction literature. Three distinct reactor models for monolith reactors
(k ⫽ 3.94 ⫻ 10⫺3 m3liquid s⫺1 kgcatalyst) in a monolith and have been published in the open literature. Hatziantoniou et al.
trickle-bed reactor with 3.5-mm impregnated catalyst particles. (1984) and Irandoust et al. (1988a) proposed models for mono-
The monolith was assumed to be operated purely in liquid lith reactors operating in the Taylor flow regime. The one
phase, the feed being saturated with hydrogen, much in excess proposed by Irandoust gained more attention and was subse-
of stoichiometric proportions. To achieve 95% conversation of quently used by other researchers (Edvinsson et al., 1994;
the limiting reactant, the simulation results showed that for a Stankiewicz et al., 2001). More recently, Liu et al. (2002) and
cell density ranging from 200 to 1100 cpsi and a washcoat Roy et al. (2002) dealt with monolith reactors operated in the
thickness of 17–70 ␮m, the required reactor volume is 6 to 53 film-flow regime, which is more akin to the trickle-flow regime
times less than that of a conventional trickle-bed reactor. The in packed beds. In the model developed by Irandoust et al.
author also compared the performance of different cell densi- (1988), and later adopted by Edvinsson et al. (1994) and others,
ties and washcoat thicknesses and found that the required plug flow was assumed in the liquid phase. Three different
monolith reactor volume increased with increasing washcoat mass transfer fluxes were considered in the differential mass
thickness and cell density. balance for reacting species, as shown in Figure 10.
Heiszwolf et al. (2001) used the conventional way of com- In all the studies involving this model, a good agreement was
paring the performance of different types of conventional re- observed between experimental studies (Irandoust et al., 1988)
actors, while studying the hydrodynamic aspects of a monolith and a superior performance compared to trickle-bed reactor
loop reactor. The authors used the power input per unit volume (TBR). However, Nijhuis et al. (2001) reported experimentally
of the system (P/V) to obtain a certain gas–liquid mass transfer determined activity of monolith per unit volume severalfold
rate as the indicator for reactor performance. The P/V for lower compared to that predicted by using the model. The
monolith reactor was calculated using the following expression authors cited reasons including irregular wall structure and
nonuniform liquid film thickness, both contributing to lower
P ⌬P L␧u L mass transfer rate in the experiment. Several other factors that
⫽ (41) could affect the performance of monolith are aeration of liquid
V L
slugs (Mewes et al., 1999), irregular liquid and gas slug length
(Sederman et al., 2003), initial maldistribution of gas and liquid
Table 8. Bed Characteristics Used in Comparing the reactants (Reinecke et al., 1996), and compression of gas slugs.
Performance of a Monolith and TBR Using Mass Transfer Important parameters for monolith reactor modeling, phase
Limited Hydrogenation of ␣-Methylstyrene over an Eggshell holdup, and contact surface area for mass transfer depend on
Nickel Catalyst* these factors. A detailed study of these hydrodynamic effects is
Monolith Trickle the current need. Carefully executed experiments and a detailed
Activity per Parameter Unit (400 cpsi) Bed computation fluid dynamic (CFD) study could help us to better
Reactor volume 3
mol/mreactor /s 6.2 4.6 understand the occurring phenomena.
Geometrical surface area 2
mol/mgeom /s 1.9 ⫻ 10⫺3 2.0 ⫻ 10⫺3 Hatziantoniou et al. (1984) modeled hydrogenation of nitro-
Catalyst amount mol/gcatalyst/s 1.1 ⫻ 10⫺5 6.2 ⫻ 10⫺5 benzoic acid (NBA) in a monolith reactor. The model was
Nickel amount mol/gnickel/s 9.8 ⫻ 10⫺4 8.4 ⫻ 10⫺5
simplified by the fact that the intrinsic reaction rate was zero
*Conditions: 373 K, 10 bar H2 (from Nijhuis et al., 2001). order with respect to NBA, and first order with respect to

2934 November 2004 Vol. 50, No. 11 AIChE Journal


Table 9. Summary of Overall Reactor Models Available in the Literature
Author Flow Regimes Basic Approach Design Equation in the Bulk Interface Equations Hydrodynamic Equations
Edvinsson
et al.,
1994
Gas–liquid slug
flow
Three distinct
mass transfer
rates, plug
冋 册 冋
d FG
dz FL
⫺NGL ⫺ NGS
⫽ N ⫺N
GL LS
册 Interface between Plug flow in both gas
and liquid slugs, no
slip between gas and
flow in each liquid slugs
phase
Refer to Figure 10 Gas/liquid

NGL,i ⫽ 共kGL a兲i 冉Pi


Hei
⫺ cL,i 冊
Liquid/solid
NLS,i ⫽ 共kLS a兲i 共cL,i ⫺ cs,i 兲

⫽ fw xcat 冘
j
␯i,j ␩j rL,j

Gas/solid
NGS,i ⫽ 共kGS a兲i 共cI,i ⫺ cs,i 兲

⫽ 共1 ⫺ fw 兲xcat 冘
j
␯i,j ␩j rI,j

Roy et al., Annular flow (gas dc Within washcoat (to calculate Gas core/liquid film
2002 core/liquid uL0 ⫽ ⫺rapp catalyst effectiveness, ␩ e (annular) regime. Both
dz
film) liquid and gas under
laminar flow with no
slip at the interface.
r app ⫽ ␥ e ␩ i (1 ⫺ OFA)r cat,v ⵜ(D e ⵜc s ) ⫽ r cat,v ; c s ⫽
c bulk at S-L interface
ⵜc s ⫽ 0 at the plane of
symmetry within the solid
wall
Within liquid
⭸c
uL,z ⫽ ⵜ共Dⵜc兲
⭸z

hydrogen. This model differs from the model proposed by in a packed-bed reactor. The authors proposed three levels of
Irandoust et al. (1988) only in that the primary mass balance of modeling. A one-dimensional, plug-flow, steady-state, convec-
hydrogen was carried out inside the washcoat as an unsteady- tion–reaction equation was advanced to describe the annular
state diffusion–reaction problem, instead of in the channel. The film. It gives the concentration profile of the reacting species
transport phenomena in the channels acted as boundary condi- along the length of the reactor. Two extra parameters, the
tions for the design equation. Unfortunately, the model also velocity profile in the film and the effectiveness factor arising
contains liquid and gas slug time constants, which cannot be out of diffusion resistance inside the catalyst, are needed to
known a priori without doing actual experiments. For this solve the above equation. They were modeled separately. To
reason, and also because of the model’s inability to incorporate obtain the effectiveness factor, the standard equation for slab
more complex reaction rate forms, it has not gained favor in geometry ␩ ⫽ tan ␾/␾ was used. The Thiele modulus ␾ ⫽
subsequent years, and therefore is not included in Table 9. l公(k/D) was found to be adequate for all shapes when the
Roy et al. (2002) published a part of their work on reactor diffusion length scale “l” was expressed as l ⫽ (1 ⫺ OFA)/
modeling of monolith reactors working in the film-flow regime. GSA. For each velocity, the velocity profile inside the film was
The authors aimed to develop criteria to determine the opti- solved using the standard Navier–Stokes equation.
mum design (washcoat thickness, channel diameter) of a The results show that for slow reactions a monolith with high
monolith based on the intrinsic kinetic rate of reactions. The solid fraction is preferred because the higher diffusion resis-
film-flow regime is achieved when the reactor is operated at tance does not have a strong bearing and more catalyst can be
very low liquid velocity (0.1 to 2 cm/s), similar to trickle flow packed in a unit volume. For fast reactions, optima exist with
respect to the diffusion length, and any further increase in
catalyst wall thickness decreases the volumetric activity.
This latest study heralds a new question in the research
pertaining to monoliths. To date, slug flow has been advocated
as being the most advantageous flow regime. A comparison of
monolith performance operating in slug-flow and film-flow
regions is in order.

Figure 10. Different mass transfer fluxes used in the Acknowledgments


monolith reactor modeling of Irandoust et al. The authors gratefully acknowledge the financial support provided by
(1989) and Edvinsson et al. (1994). Bayer Technology Services, Bayer AG, Germany.

AIChE Journal November 2004 Vol. 50, No. 11 2935


Notation b ⫽ bubble
c ⫽ capillary, channel
a ⫽ interfacial area per unit volume, 1/L D ⫽ Darcy
C ⫽ factor F ⫽ Fanning
C ⫽ distribution parameter fr, f ⫽ frictional
cpsi ⫽ cells (or channels) per square inch g ⫽ gas
D ⫽ diffusion coefficient, m2/s G ⫽ gas
db ⫽ bubble diameter GL ⫽ gas–liquid
dc ⫽ capillary diameter GS ⫽ gas–solid
dh ⫽ hydraulic diameter m h ⫽ hydraulic
dP ⫽ dispersed bubble diameter, m l ⫽ liquid
dp/dz ⫽ pressure gradient, Pa/m L ⫽ liquid
DR ⫽ reactor diameter, m LF ⫽ liquid film alongside a Taylor bubble
F ⫽ factor LS ⫽ liquid–solid
f ⫽ friction factor m ⫽ measured
g ⫽ gravity, m/s2 orif ⫽ orifice
GSA ⫽ geometric surface area, in.2/in.3 P ⫽ dispersed bubble
k ⫽ mass transfer coefficient, m/s R ⫽ reactor
L ⫽ length, m s ⫽ slug
L ⫽ cell spacing, in. st ⫽ hydrostatic
Lb ⫽ gas bubble length, m tot ⫽ total
Lb ⫽ constant bubble length (dc ⫽ 2 mm), m TP ⫽ two phase
Lc ⫽ channel length z ⫽ length
LR ⫽ reactor length, m
Ls ⫽ liquid slug length, m
MR ⫽ monolith reactor Superscripts
n ⫽ cell density, cpsi
OFA ⫽ open frontal area * ⫽ dimensionless velocity
P/V ⫽ power input per unit volume of the system, W/m3 n ⫽ scaling factor
R ⫽ radius of the capillary, m
R ⫽ radius of gas plug, m
Re ⫽ Reynolds number ⫽ vd␳/␮ Literature Cited
Sc ⫽ Schmidt number ⫽ ␮␳/D
Sh ⫽ Sherwood number ⫽ kd/D Albers, R. E., M. Nyström, M. Siverström, A. Sellin, A. C. Dellve, U.
t ⫽ wall thickness, mm Andersson, H. Herrmann, and Th. Berglin, “Development of Monolith
TBR ⫽ trickle-bed reactor Based Process for H2O2 Production: From Idea to Large-Scale Imple-
UCL ⫽ unit cell length ⫽ Lb ⫹ Ls mentation,” Catal. Today, 69, 247 (2001).
uG ⫽ superficial gas velocity, m/s Al-Dahhan, M. H., and W. Highfill, “Liquid Holdup Measurement Tech-
uL ⫽ superficial liquid velocity, m/s niques in Laboratory High Pressure Trickle Bed Reactors,” Can.
uL,m ⫽ maximum liquid velocity, m/s J. Chem. Eng., 77(4), 759 (1999).
utot ⫽ sum of superficial gas and liquid velocity (uL ⫹ uG), m/s Al-Dahhan, M. H., F. Larachi, M. P. Dudukovic, and A. Laurent, “High-
V ⫽ volume, m3 Pressure Trickle-Bed Reactors: A Review,” Ind. Eng. Chem. Res., 36(8),
v ⫽ cell velocity, m/s 3292 (1997).
vb ⫽ bubble velocity, m/s Andersson, B., S. Irandoust, and A. Cybulski, “Modeling of Monolith
vG ⫽ gas velocity, m/s Reactors in Three-Phase Processes,” Chem. Ind. Struct. Catal. & Reac-
VG,u ⫽ drift velocity, m/s tors (Dekker), 71, 267 (1998).
We ⫽ Weber number ⫽ ␳lv 2l dp/␴ Bercić, G., and A. Pintar, “The Role of Gas Bubbles and Liquid Slug
x ⫽ quality Lengths on Mass Transport in the Taylor Flow through Capillaries,”
⌬p ⫽ pressure drop Chem. Eng. Sci., 52(21–22), 3709 (1997).
⌬P/⌬L ⫽ pressure drop per unit length, Pa/m Biswas, J., and P. F. Greenfield, “Two-Phase Flow through Vertical Cap-
⌬PG ⫽ pressure drop caused by gas, Pa illaries—Existence of a Stratified Flow Pattern,” Int. J. Multiphase Flow,
11(4), 553 (1985).
Boyer, C., and B. Fanget, “Measurement of Liquid Flow Distribution in
Greek letters Trickle Bed Reactor of Large Diameter with a New Gamma-Ray To-
mographic System,” Chem. Eng. Sci., 57, 1079 (2002).
␣ ⫽ dimensionless film thickness ⫽ ␦f/dt Broekhuis, R. R., R. M. Machado, and A. F. Nordquist, “The Ejector-
⌽ ⫽ volumetric flow rate, m3/s Driven Monolith Loop Reactor—Experiments and Modeling,” Catal.
⌿s ⫽ liquid slug length, m Today, 69(1– 4), 87 (2001).
␦f ⫽ liquid film thickness, m Chen, J., R. Novica, M. H. Al-Dahhan, and M. P. Dudukovic, “Study of
␦f,m ⫽ maximum film thickness, m Particle Motion in Packed/Ebullated Beds by Computed Tomography
␧ ⫽ bed void fraction (CT) and Computer Automated Radioactive Particle Tracking
␧G , ␤ ⫽ gas holdup (CARPT),” AIChE J., 47(5), 994 (2001).
␧gG ⫽ volume fraction of bubbles Coleman, J. W., and S. Garimella, “Characterization of Two-Phase Flow
␧gG/Lb ⫽ number of bubbles per unit length Patterns in Small Diameter Round and Rectangular Tubes,” Int. J. Heat
␧L ⫽ liquid holdup Mass Transfer, 42(15), 2869 (1999).
␥ ⫽ Lc/dc Crynes, L. L., R. L. Cerro, and M. A. Abraham, “Monolith Froth Reactor:
␮ ⫽ dynamic viscosity, kg/ms Development of a Novel Three-Phase Catalytic System,” AIChE J.,
␳ ⫽ density, kg/m3 41(2), 337– 45 (1995).
␴ ⫽ surface tension, N/m Cybulski, A., R. Edvinsson, S. Irandoust, and B. Andersson, “Liquid-Phase
␨ ⫽ dimensionless tube length Methanol Synthesis: Modelling of a Monolithic Reactor,” Chem. Eng.
Sci., 48(20), 3463 (1993).
Subscripts Cybulski, A., and J. A. Moulijn, Structured Catalysts and Reactors, Marcel
Dekker, New York (1998).
A ⫽ species A Cybulski, A., A. Stankiewicz, R. K. Edvinsson Albers, and J. A. Moulijn,
B ⫽ species B “Monolithic Reactors for Fine Chemicals Industries: A Comparative

2936 November 2004 Vol. 50, No. 11 AIChE Journal


Analysis of a Monolithic Reactor and a Mechanically Agitated Slurry Taylor Flow through a Capillary,” Can. J. Chem. Eng., 70(1), 115
Reactor,” Chem. Eng. Sci., 54(13–14), 2351 (1999). (1992).
Edvinsson, R. K., and A. Cybulski, “A Comparative Analysis of the Irandoust, S., and O. Gahne, “Competitive Hydrodesulfurization and Hy-
Trickle-Bed and the Monolithic Reactor for Three-Phase Hydrogena- drogenation in a Monolithic Reactor,” AIChE J., 36(5), 746 (1990).
tions,” Chem. Eng. Sci., 49(24), 5653 (1994). Kapteijn, F., T. A. Nijhuis, J. J. Heiszwolf, and J. A. Moulijn, “New
Edvinsson, R. K., A. M. Holmgren, and S. Irandoust, “Liquid-Phase Non-Traditional Multiphase Catalytic Reactors Based on Monolithic
Hydrogenation of Acetylene in a Monolithic Catalyst Reactor,” Ind. Structures,” Catal. Today, 66(2– 4), 133 (2001).
Eng. Chem. Res., 34(1), 94 (1995). Kawakami, K., K. Kawasaki, F. Shiraishi, and K. Kusunoki, “Performance
Edvinsson, R. K., M. J. J. Houterman, T. Vergunst, E. Grolman, and J. A. of a Honeycomb Monolith Bioreactor in a Gas–Liquid–Solid Three-
Moulijn, “Novel Monolithic Stirred Reactor,” AIChE J., 44(11), 2459 Phase System,” Ind. Eng. Chem. Res., 28(4), 394 (1989).
(1998). Klinghoffer, A. A., R. L. Cerro, and M. A. Abraham, “Catalytic Wet
Edvinsson, R. K., and S. Irandoust, “Hydrodesulfurization of Dibenzothio- Oxidation of Acetic Acid Using Platinum on Alumina Monolith Cata-
phene in a Monolithic Catalyst Reactor,” Ind. Eng. Chem. Res., 32(2), lyst,” Catal. Today, 40(1), 59 (1998a).
391 (1993). Klinghoffer, A. A., R. L. Cerro, and M. A. Abraham, “Influence of Flow
Ellenberger, J., and R. Krishna, “Counter-Current Operation of Structured Properties on the Performance of the Monolith Froth Reactor for Cata-
Catalytically Packed Distillation Columns: Pressure Drop, Holdup and lytic Wet Oxidation of Acetic Acid,” Ind. Eng. Chem. Res., 37(4), 1203
Mixing,” Chem. Eng. Sci., 54(10), 1339 (1999). (1998b).
Folger, H. S., Elements of Chemical Reaction Engineering, Prentice Hall, Kolb, W. B., and R. L. Cerro, “Coating the Inside of a Capillary of Square
Englewood Cliffs, NJ (1992). Cross Section,” Chem. Eng. Sci., 46(9), 2181 (1991).
Froment, G. B., and K. B. Bischoff, Chemical Reactor Analysis and Kolb, W. B., and R. L. Cerro, “Film Flow in the Space between a Circular
Design, Wiley, New York (1990). Bubble and a Square Tube,” J. Colloid Interface Sci., 159(2), 302
Garcia-Bordeje, E., F. Kapteijn, and J. A. Moulijn, “Preparation and (1993).
Characterisation of Carbon-Coated Monoliths for Catalyst Supports,” Kreutzer, M. T., P. Du, J. J. Heiszwolf, F. Kapteijn, and J. A. Moulijn,
Carbon, 40(7), 1079 (2002). “Mass Transfer Characteristics of Three-Phase Monolith Reactors,”
Gulati, S. T., “Ceramic Catalyst Supports for Gasoline Fuel,” Structured Chem. Eng. Sci., 56(21–22), 6015 (2001).
Catalysts and Reactors, A. Cybulski and J. A. Moulijn, eds., Marcel Krishna, R., and S. T. Sie, “Strategies for Multiphase Reactor Selection,”
Dekker, New York, pp. 15–58 (1998). Chem. Eng. Sci., 49(24A), 4029 (1994).
Harter, I., C. Boyer, L. Raynal, G. Ferschneider, and T. Gauthier, “Flow Kumar, S. B., M. P. Dudukovic, J. Chaouki, F. Larachi, and M. P.
Distribution Studies Applied to Deep Hydro-Desulfurization,” Ind. Eng. Dudukovic, “Computer Assisted Gamma and X-ray Tomography: Ap-
Chem. Res., 40(23), 5262 (2001). plications to Multiphase Flow Systems,” Non-Invasive Monitoring of
Hatziantoniou, V., and B. Andersson, “Solid–Liquid Mass Transfer in Multiphase Flows, Elsevier, Amsterdam, pp. 47–103 (1997).
Segmented Gas–Liquid Flow through a Capillary,” Ind. Eng. Chem. Lachman, I. M., and J. L. Williams, “Extruded Monolithic Catalyst Sup-
ports,” Catal. Today, 14(2), 317 (1992).
Fundam., 21(4), 451 (1982).
Lebens, P. J. M., J. J. Heiszwolf, F. Kapteijn, S. T. Sie, and J. A. Moulijn,
Hatziantoniou, V., and B. Andersson, “The Segmented Two-Phase Flow
“Gas–Liquid Mass Transfer in an Internally Finned Monolith Operated
Monolithic Catalyst Reactor. An Alternative for Liquid-Phase Hydro-
Countercurrently in the Film Flow Regime,” Chem. Eng. Sci., 54(21),
genations,” Ind. Eng. Chem. Fundam., 23(1), 82 (1984).
5119 (1999a).
Hatziantoniou, V., B. Andersson, and N. H. Schoon, “Mass Transfer and
Lebens, P. J. M., F. Kapteijn, S. T. Sie, and J. A. Moulijn, “Potentials of
Selectivity in Liquid-Phase Hydrogenation of Nitro Compounds in a
Internally Finned Monoliths as a Packing for Multifunctional Reactors,”
Monolithic Catalyst Reactor with Segmented Gas–Liquid Flow,” Ind.
Chem. Eng. Sci., 54(10), 1359 (1999b).
Eng. Chem. Proc. Des. Dev., 25(4), 964 (1986).
Lebens, P. J. M., M. M. Stork, F. Kapteijn, S. T. Sie, and J. A. Moulijn,
Heibel, A. K., J. J. Heiszwolf, F. Kapteijn, and J. A. Moulijn, “Influence of
“Hydrodynamics and Mass Transfer Issues in a Countercurrent Gas–
Channel Geometry on Hydrodynamics and Mass Transfer in the Mono- Liquid Internally Finned Monolith Reactor,” Chem. Eng. Sci., 54(13–
lith Film Flow Reactor,” Catal. Today, 69(1– 4), 153 (2001a). 14), 2381 (1999c).
Heibel, A. K., T. W. J. Scheenen, J. J. Heiszwolf, H. Van As, F. Kapteijn, Lebens, P. J. M., R. van der Meijden, R. K. Edvinsson, F. Kapteijn, S. T.
and J. A. Moulijn, “Gas and Liquid Phase Distribution and Their Effect Sie, and J. A. Moulijn, “Hydrodynamics of Gas–Liquid Countercurrent
on Reactor Performance in the Monolith Film Flow Reactor,” Chem. Flow in Internally Finned Monolithic Structures,” Chem. Eng. Sci.,
Eng. Sci., 56(21–22), 5935 (2001b). 52(21–22), 3893 (1997).
Heiszwolf, J. J., B. L. Engelvaart, G. M. van den Eijnden, T. M. Kreutzer, Levenspiel, O., The Chemical Reactor Omnibook, OSU Book Stores, Inc.,
F. Kapteijn, and J. A. Moulijn, “Hydrodynamic Aspects of the Monolith Corvallis, OR (1996).
Loop Reactor,” Chem. Eng. Sci., 56(3), 805 (2001a). Levenspiel, O., Chemical Reaction Engineering, Wiley, New York (1998).
Heiszwolf, J. J., M. T. Kreutzer, M. G. van den Eijnden, F. Kapteijn, and Liu, W., “Ministructured Catalyst Bed for Gas–Liquid–Solid Multiphase
J. A. Moulijn, “Gas–Liquid Mass Transfer of Aqueous Taylor Flow in Catalytic Reaction,” AIChE J., 48(7), 1519 (2002).
Monoliths,” Catal. Today, 69(1– 4), 51 (2001b). Liu, W., W. P. Addiego, C. M. Sorensen, and T. Boger, “Monolith Reactor
Highfill, W., and M. H. Al-Dahhan, “Liquid–Solid Mass Transfer Coeffi- for the Dehydrogenation of Ethylbenzene to Styrene,” Ind. Eng. Chem.
cient in High Pressure Trickle Bed Reactors,” Trans. IChemE, 79(A) Res., 41(13), 3131 (2002).
(2001). Marcandelli, C., A. S. Lamine, J. R. Bernard, and G. Wild, “Liquid
Iliuta, I., and F. Larachi, “Mechanistic Model for Structured-Packing Distribution in Trickle-Bed Reactor,” Oil Gas Sci Technol. Rev. IFP,
Containing Columns Irrigated Pressure Drop, Liquid Holdup, and Pack- 55(4), 407 (2000).
ing Fractional Wetted Area,” Ind. Eng. Chem. Res., 40, 5140 (2000). Mazzaroni, I., and G. Baldi, “Liquid-Phase Hydrogenation on a Monolith
Irandoust, S., and B. Andersson, “Mass Transfer and Liquid-Phase Reac- Catalyst,” Recent Trends in Chemical Reaction Engineering, Vol. 2,
tions in a Segmented Two-Phase Flow Monolithic Catalyst Reactor,” B. D. Kulkarni, R.A. Mashelkar, and M. M. Sharma, eds., Wiley Eastern
Chem. Eng. Sci., 43(8), 1983 (1988a). Ltd. New Delhi, p. 181 (1987).
Irandoust, S., and B. Andersson, “Monolithic Catalysts for Nonautomobile Mewes, D., T. Loser, and M. Millies, “Modelling of Two-Phase Flow in
Applications,” Catal. Rev. Sci. Eng., 30(3), 341 (1988b). Packings and Monoliths,” Chem. Eng. Sci., 54(21), 4729 (1999).
Irandoust, S., and B. Andersson, “Liquid Film in Taylor Flow through a Mishima, K., and T. Hibiki, “Some Characteristics of Air–Water Two-
Capillary,” Ind. Eng. Chem. Res., 28(11), 1684 (1989a). Phase Flow in Small Diameter Vertical Tubes,” Int. J. Multiphase Flow,
Irandoust, S., and B. Andersson, “Simulation of Flow and Mass Transfer in 22(4), 703 (1996).
Taylor Flow through a Capillary,” Comput. Chem. Eng., 13(4 –5), 519 Mishima, K., and M. Ishii, “Flow Regime Transition Criteria for Upward
(1989b). Two-Phase Flow in Vertical Tubes,” Int. J. Heat Mass Transfer, 27(5),
Irandoust, S., B. Andersson, E. Bengtsson, and M. Siverstroem, “Scaling 723 (1984).
Up of a Monolithic Catalyst Reactor with Two-Phase Flow,” Ind. Eng. Mishra, V. S., V. V. Mahajani, and J. B. Joshi, “Wet Air Oxidation,” Ind.
Chem. Res., 28(10), 1489 (1989c). Eng. Chem. Res., 34(1), 2 (1995).
Irandoust, S., S. Ertle, and B. Andersson, “Gas–Liquid Mass Transfer in Nemec, D., G. Bercić, and J. Levec, “Gravimetric Method for the Deter-

AIChE Journal November 2004 Vol. 50, No. 11 2937


mination of Liquid Holdup in Pressurized Trickle-Bed Reactors,” Ind. Aspects from Microkinetics to Reactor Design,” Proc. of the 6th Natl.
Eng. Chem. Res., 40(15), 3418 (2001). Sympos. Chem. React. Eng., Warsaw, Poland (1989).
Nijhuis, T. A., F. M. Dautzenberg, and J. A. Moulijn, “Modeling of Schutt, B. D., B. Serrano, R. L. Cerro, and M. A. Abraham, “Production of
Monolithic and Trickle-Bed Reactors for the Hydrogenation of Styrene,” Chemicals from Cellulose and Biomass-Derived Compounds through
Chem. Eng. Sci., 58, 1113 (2003). Catalytic Sub-Critical Water Oxidation in a Monolith Reactor,” Biomass
Nijhuis, T. A., M. T. Kreutzer, A. C. J. Romijn, F. Kapteijn, and J. A. & Bioenergy, 22(5), 365 (2002).
Moulijn, “Monolithic Catalysts as Efficient Three-Phase Reactors,” Smits, H. A., A. Stankiewicz, W. C. Glasz, T. H. A. Fogl, and J. A.
Chem. Eng. Sci., 56(3), 823 (2001a). Moulijn, “Selective Three-Phase Hydrogenation of Unsaturated Hydro-
Nijhuis, T. A., M. T. Kreutzer, A. C. J. Romijn, F. Kapteijn, and J. A. carbons in a Monolithic Reactor,” Chem. Eng. Sci., 51(11), 3019 (1996).
Moulijn, “Monolithic Catalysts as More Efficient Three-Phase Reac- Stankiewicz, A., “Process Intensification in In-Line Monolithic Reactor,”
tors,” Catal. Today, 66(2– 4), 157 (2001b). Chem. Eng. Sci., 56(2), 359 (2001).
Patrick, R. H., Jr., T. Klindera, L. L. Crynes, R. L. Cerro, and M. A. Stankiewicz, A. I., and J. A. Moulijn, “Process Intensification: Transform-
Abraham, “Residence Time Distribution in Three-Phase Monolith Re- ing Chemical Engineering,” Chem. Eng. Prog., 96(1), 22 (2000).
actor,” AIChE J., 41(3), 649 (1995). Thulasidas, T. C., M. A. Abraham, and R. L. Cerro, “Bubble-Train Flow in
Patrick, T. A., and M. A. Abraham, “Evaluation of a Monolith-Supported Capillaries of Circular and Square Cross Section,” Chem. Eng. Sci.,
Pt/Al2O3 Catalyst for Wet Oxidation of Carbohydrate-Containing Waste 50(2), 183 (1995a).
Streams,” Environ. Sci. Technol., 34, 3480 (2000). Thulasidas, T. C., M. A. Abraham, and R. L. Cerro, “Dispersion during
Pinna, D., E. Tronconi, and L. Tagliabue, “High Interaction Regime Bubble-Train Flow in Capillaries,” Chem. Eng. Sci., 54(1), 61 (1999).
Lockhart–Martinelli Model for Pressure Drop in Trickle-Bed Reactors,” Thulasidas, T. C., R. L. Cerro, and M. A. Abraham, “The Monolith Froth
AIChE J., 47(1), 19 (2001). Reactor: Residence Time Modeling and Analysis,” Chem. Eng. Res.
Ramachandran, P. A., and R. V. Chaudhari, Three-Phase Catalytic Reac- Des., 73(A3), 314 (1995b).
tors, Gordon & Breach, New York (1983). Triplett, K. A., S. M. Ghiaasiaan, S. I. Abdel-Khalik, A. LeMouel, and
Reinecke, N., and D. Mewes, “Flow Regimes of Two Phase Flow in B. N. McCord, “Gas–Liquid Two-Phase Flow in Microchannels. Part II:
Monolith Catalyst,” Proc. of the 5th World Congress on Chemical Void Fraction and Pressure Drop,” Int. J. Multiphase Flow, 25(3), 395
Engineering, San Diego, CA (1996). (1999a).
Reinecke, N., and D. Mewes, “Oscillatory Transient Two-Phase Flows in Triplett, K. A., S. M. Ghiaasiaan, S. I. Abdel-Khalik, and D. L. Sadowski,
Single Channels with Reference to Monolithic Catalyst Supports,” Int. J. “Gas–Liquid Two-Phase Flow in Microchannels. Part I: Two-Phase
Multiphase Flow, 25(6 –7), 1373 (1999). Flow Patterns,” Int. J. Multiphase Flow, 25(3), 377 (1999b).
Reinecke, N., G. Petritsch, D. Schmitz, and D. Mewes, “Tomographic Van’t Riet, K., “Review of Measuring Methods and Results in Nonviscous
Measurement Techniques—Visualization of Multiphase Flows,” Chem. Gas–Liquid Mass Transfer in Stirred Vessels,” Ind. Eng. Chem. Proc.
Eng. Technol., 21, 7 (1998). Dev., 18, 357 (1979).
Roy, S., A. K. Heibel, W. Liu, and T. Boger, “Design of Monolithic Williams, J. L., “Monolith Structures, Materials, Properties and Uses,”
Catalysts for Multiphase Reactions,” Oral Presentation, 17th SCRE Catal. Today, 69(1– 4), 3 (2001).
(2002). Woehl, P., and R. L. Cerro, “Pressure Drop in Monolith Reactors,” Catal.
Satterfield, C. N., and F. Ozel, “Some Characteristics of Two-Phase Flow Today, 69(1– 4), 171 (2001).
in Monolithic Catalyst Structures,” Ind. Eng. Chem. Fundam., 16(1), 61 Zhao, T. S., and Q. C. Bi, “Co-Current Air–Water Two-Phase Flow
(1977). Patterns in Vertical Triangular Microchannels,” Int. J. Multiphase Flow,
Schluter, V., and W. D. Deckwer, “Gas–Liquid Mass Transfer in Stirred 27(5), 765 (2001).
Vessels,” Chem. Eng. Sci., 47, 2357 (1992).
Schöön, N. H., “Recent Progress in Liquid-Phase Hydrogenation: With Manuscript received July 14, 2003, and revision received Feb. 29, 2004.

2938 November 2004 Vol. 50, No. 11 AIChE Journal

You might also like